This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

KEK-TH-2358
Gauge Symmetry Restoration by Higgs Condensation
in Flux Compactifications on Coset Spaces

Satoshi Isoa,b , Noriaki Kitazawac , Takao Suyamaa

a Theory Center, High Energy Accelerator Research Organization (KEK),
Tsukuba, Ibaraki 305-0801, Japan
bGraduate University for Advanced Studies (SOKENDAI),
Tsukuba, Ibaraki 305-0801, Japan
cDepartment of Physics, Tokyo Metropolitan University,
Hachioji, Tokyo 192-0397, Japan
E-mail: iso(at)post.kek.jpE-mail: noriaki.kitazawa(at)tmu.ac.jpE-mail: tsuyama(at)post.kek.jp
Abstract

Extra-dimensional components of gauge fields in higher-dimensional gauge theories will play a role of the Higgs field and become tachyonic after Kaluza-Klein compactifications on internal spaces with (topologically nontrivial) gauge field backgrounds. Its condensation is then expected to break gauge symmetries spontaneously. But, contrary to the expectation, some models exhibit restoration of gauge symmetries. In this paper, by considering all the massive Kaluza-Klein excitations of gauge fields, we explicitly show that some of them indeed become massless at the minimum of the Higgs potential and restore (a part of) the gauge symmetries which are broken by gauge field backgrounds. We particularly consider compactifications on S2S^{2} with monopole-like fluxes and also on 2\mathbb{CP}^{2} with instanton and monopole-like fluxes. In some cases, the gauge symmetry is fully restored, as argued in previous literatures. In other cases, there is a stable vacuum with a partial restoration of the gauge symmetry after Higgs condensation. Topological structure of the gauge field configurations prevent the gauge symmetries to be restored.

1 Introduction

The dynamics of gauge symmetry breaking is yet to be investigated, especially when it is caused by the elementary Higgs scalar field with a non-trivial potential. The mechanism of gauge symmetry breaking or the origin of the Higgs potential is highly required. Among numerous proposals or models including radiative symmetry breaking mechanism (see e.g., [1, 2, 3, 4, 6, 5, 7]) and extra dimensions (see e.g., [8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24]), a possibility to understand the origin of Higgs potential in the context of higher dimensional gauge theory has been widely investigated. In the present paper, we revisit the gauge symmetry breaking by the coset space dimensional reductions of higher-dimensional gauge theories with background gauge fluxes (see [25] for review, and see e.g., [26] for dimensional reduction to non-coset spaces). The basic idea of this construction appeared in [8] which realizes the bosonic part of the Weinberg-Salam model from the six-dimensional Yang-Mills theory. In this construction, the Higgs potential of the double-well type dynamically appears, and this class of models have been applied to the gauge-Higgs unification models of the electroweak theory [20].

Background gauge fluxes in compact spaces in higher-dimensional gauge theories are originally introduced to stabilize the compact space [27, 28, 29, 30, 31, 32] in the context of the Einstein-Yang-Mills theory, and further developed in the studies of flux compactifications in string theories (for reviews, see [33, 34, 35] ). The well studied examples of compact spaces are coset spaces G/HG/H, such as S2=S^{2}=SU(2)/(2)/U(1)(1) or 2=\mathbb{CP}^{2}=SU(3)/((3)/(SU(2)×(2)\timesU(1)(1)), and (in)stability of such compactifications in the presence of gravity have been extensively investigated. If tachyonic fields appear, the solution becomes unstable and their condensations will generate a new vacuum solution. In particular, such tachyonic fields can be utilized as candidates of the Higgs scalars, and understanding of the shape of tachyon potential and the pattern of gauge symmetry breaking is an important issue to be investigated.

In flux compactifications, the original gauge symmetry in higher-dimensions is explicitly broken by the background gauge fluxes in the compact spaces, and the Higgs vacuum expectation value is expected to further break some part of the remaining gauge symmetries spontaneously in four-dimensional effective theory. In previous literatures, most studies have been focussed on low lying states in the effective four-dimensional theories after compactifications. Among an infinitely many fields, only massless fields are usually taken into considerations in the effective theory, and all the other higher excited modes are neglected. It is justified when we consider low-energy physics below the scale of the compact spaces, but when we investigate the condensation of the tachyonic field, the massive modes will also play an important role since the mass scale in the Higgs potential is typically the same as masses of Kaluza-Klein higher modes. Especially, when the Higgs acquires vacuum expectation value, we need to take care of a possibility that some of the massive Kaluza-Klein modes may become massless.

In this paper, we investigate dynamics of the Higgs condensation in several simple models on coset spaces, such as S2=S^{2}=SU(2)/(2)/U(1)(1) and 2=\mathbb{CP}^{2}=SU(3)/((3)/(SU(2)×(2)\timesU(1)(1)), with all the massive Kaluza-Klein modes included. We find that, although the Higgs vacuum expectation value itself breaks a part of the remaining gauge symmetries and the corresponding gauge bosons indeed become massive, some of the massive Kaluza-Klein modes will become massless and gauge symmetries are recovered in the four-dimensional effective theory. In fact, such a possibility was pointed out in [36, 37]. In this paper, we develop a group-theoretic technique which enables us to clarify explicitly which Kaluza-Klein modes become massless vectors after the Higgs condensation. In string theory, it is known that similar gauge symmetry enhancement occurs for condensation of massless scalars, or moduli. Non-supersymmetric string theories are discussed in this context recently in [38] and references therein.

In section 2, we give a general formulation of the Kaluza-Klein reduction on coset spaces G/HG/H with a topologically non-trivial background gauge field configuration. In section 3, we introduce the notion of “symmetric field” [39] which corresponds to the zero mode, or a constant mode on the flat compact space without flux. Interestingly, some of these symmetric fields may have non-trivial potentials with a negative mass squared at the origin, and we call them symmetric Higgs fields. In section 4, we investigate the Higgs condensation in gauge field theories compactified on S2S^{2}. We study three different types of models, whose background monopole-like fluxes are different. We particularly investigate patterns of gauge symmetry breaking when the symmetric Higgs fields have vacuum expectation value. In section 5 we generalize the analysis on S2S^{2} to 2\mathbb{CP}^{2} coset models. In this case, since the coset space is SU(3)/(SU(2)×U(1))\rm SU(3)/(SU(2)\times U(1)), both of instanton and monopole-like background configurations can exist. In one of the examples we study, all the gauge symmetries are restored by the Higgs condensation, which cancels the background gauge flux as was pointed out in [36, 37]. There exists another type of models, in which a topologically nontrivial gauge field fluxes prevent the gauge symmetries to be recovered, and a stable vacuum with a partial restoration of gauge symmetries is realized. In the last section we summarize our results and conclude.

There are several Apendices which review various materials necessary for our investigations. In Appendix A, we review the basics of coset space G/HG/H, and describe GG as a principal HH-bundle in Appendix B. In Appendix C, we review the construction of the background gauge field and the vielbein on G/HG/H which are provided by the Maurer-Cartan 1-form on GG. In Appendix D, we review a proof that the background gauge field satisfies the equations of motion. In Appendix E, concrete forms of the background gauge field and the vielbein are given in the case of S2=SU(2)/U(1)S^{2}={\rm SU}(2)/{\rm U}(1). In Appendix F, we review a construction of mode expansions on G/HG/H by using the Peter-Weyl theorem for the mode expansions on GG. In Appendix G, we explain eigenvalues of the Laplacian on mode functions and mass formula of various fields on G/HG/H. We also show that the symmetric Higgs field has a negative mass squared and becomes tachyonic. In Appendix H, we prove that the symmetric Higgs field satisfies the condition of the symmetric field on G/HG/H.

2 Kaluza-Klein reduction on coset spaces

2.1 Action in background gauge fields

We consider Yang-Mills theory on a (4+d)(4+d)-dimensional manifold 4×\mathbb{R}^{4}\times{\cal M} with the action

S=𝑑vTr[14FMNFMN],dv:=1gYM2d4+dXG,S\ =\ \int dv\,{\rm Tr}\left[-\frac{1}{4}F_{MN}F^{MN}\right],\hskip 28.45274ptdv\ :=\ \frac{1}{g_{\rm YM}^{2}}d^{4+d}X\sqrt{-G}, (2.1)

where M,N=0,1,,3+dM,N=0,1,\cdots,3+d and GMNG_{MN} is a metric on 4×\mathbb{R}^{4}\times{\cal M}. The overall normalization of the action is chosen such that each matrix component of the gauge field AMA_{M} is canonically normalized. Our convention for the field strength is

FMN:=MANNAM+i[AM,AN],F_{MN}\ :=\ \nabla_{M}A_{N}-\nabla_{N}A_{M}+i[A_{M},A_{N}], (2.2)

where M\nabla_{M} is the covariant derivative with respect to the metric GMNG_{MN}.

We investigate this theory around a background gauge field A¯M\bar{A}_{M}. The gauge field AMA_{M} is then decomposed as AM=A¯M+aMA_{M}=\bar{A}_{M}+a_{M}. In the following, we often use the notation

D¯MaN:=MaN+i[A¯M,aN].\bar{D}_{M}a_{N}\ :=\ \nabla_{M}a_{N}+i[\bar{A}_{M},a_{N}]. (2.3)

We employ the background field gauge

D¯MaM=MaM+i[A¯M,aM]= 0.\bar{D}^{M}a_{M}\ =\ \nabla^{M}a_{M}+i[\bar{A}^{M},a_{M}]\ =\ 0. (2.4)

The corresponding gauge-fixing term is given by

Sgf=𝑑vTr[12(D¯MaM)2].S_{\rm gf}\ =\ \int dv\,{\rm Tr}\left[-\frac{1}{2}\left(\bar{D}^{M}a_{M}\right)^{2}\right]. (2.5)

By expanding aMa_{M} into Kaluza-Klein modes on {\cal M}, we can obtain a four-dimensional gauge theory coupled to various matter fields.

Let xμx^{\mu} (μ=0,,3\mu=0,\cdots,3) be coordinates on 4\mathbb{R}^{4}, and let yαy^{\alpha} (α=1,,d\alpha=1,\cdots,d) be coordinates on \cal M. Accordingly, the gauge field aMa_{M} is decomposed into aμa_{\mu} and ϕα\phi_{\alpha}. We assume that the background gauge field A¯M\bar{A}_{M} is of the form

A¯M=(0,A¯α),μA¯α= 0.\bar{A}_{M}\ =\ (0,\bar{A}_{\alpha}),\hskip 28.45274pt\partial_{\mu}\bar{A}_{\alpha}\ =\ 0. (2.6)

This means that we put an xμx^{\mu}-independent gauge flux on {\cal M}. Note that the extra-dimensional components of the gauge field, ϕα\phi_{\alpha}, provide a set of adjoint matters, transforming homogeneously under the gauge transformations as they are defined by a difference of two gauge fields AαA_{\alpha} and A¯α\bar{A}_{\alpha}. We also set the background metric of 4×\mathbb{R}^{4}\times{\cal M} as

GMN=[ημν00hαβ(y)].G_{MN}\ =\ \left[\begin{array}[]{cc}\eta_{\mu\nu}&0\\ 0&h_{\alpha\beta}(y)\end{array}\right]. (2.7)

The total action S+SgfS+S_{\rm gf} consists of the following three parts:

S1\displaystyle S_{1} =\displaystyle= 𝑑vTr[14FμνFμν12(μaμ)2]\displaystyle\int dv\,{\rm Tr}\left[-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}-\frac{1}{2}\left(\partial^{\mu}a_{\mu}\right)^{2}\right] (2.8)
=\displaystyle= 𝑑vTr[12(μaν)2iμaν[aμ,aν]+14[aμ,aν]2],\displaystyle\int dv\,{\rm Tr}\left[-\frac{1}{2}\left(\partial_{\mu}a_{\nu}\right)^{2}-i\partial_{\mu}a_{\nu}[a^{\mu},a^{\nu}]+\frac{1}{4}[a_{\mu},a_{\nu}]^{2}\right],
S2\displaystyle S_{2} =\displaystyle= 𝑑vTr[14FμαFμαμaμD¯αϕα]\displaystyle\int dv\,{\rm Tr}\left[-\frac{1}{4}F_{\mu\alpha}F^{\mu\alpha}-\partial^{\mu}a_{\mu}\bar{D}^{\alpha}\phi_{\alpha}\right] (2.9)
=\displaystyle= 𝑑vTr[12(Dμϕα)212(D¯αaμ)2+i[aμ,ϕα]D¯αaμ],\displaystyle\int dv\,{\rm Tr}\left[-\frac{1}{2}\left(D_{\mu}\phi_{\alpha}\right)^{2}-\frac{1}{2}\left(\bar{D}_{\alpha}a_{\mu}\right)^{2}+i[a_{\mu},\phi_{\alpha}]\bar{D}^{\alpha}a^{\mu}\right],

where

Dμϕα:=μϕα+i[aμ,ϕα],D_{\mu}\phi_{\alpha}\ :=\ \partial_{\mu}\phi_{\alpha}+i[a_{\mu},\phi_{\alpha}], (2.10)

and

S3\displaystyle S_{3} =\displaystyle= 𝑑vTr[14FαβFαβ12(D¯αϕα)2]\displaystyle\int dv\,{\rm Tr}\left[-\frac{1}{4}F_{\alpha\beta}F^{\alpha\beta}-\frac{1}{2}\left(\bar{D}^{\alpha}\phi_{\alpha}\right)^{2}\right]
=\displaystyle= 𝑑vTr[14(F¯αβ+D¯αϕβD¯βϕα+i[ϕα,ϕβ])212(D¯αϕα)2],\displaystyle\int dv\,{\rm Tr}\left[-\frac{1}{4}\left(\bar{F}_{\alpha\beta}+\bar{D}_{\alpha}\phi_{\beta}-\bar{D}_{\beta}\phi_{\alpha}+i[\phi_{\alpha},\phi_{\beta}]\right)^{2}-\frac{1}{2}\left(\bar{D}^{\alpha}\phi_{\alpha}\right)^{2}\right],

where F¯αβ\bar{F}_{\alpha\beta} is the background field strength of the gauge potential A¯α\bar{A}_{\alpha}.

In the Kaluza-Klein reduction, the terms (LABEL:scalar_potential_original) in S3S_{3} give the scalar potential V(ϕ)V(\phi) after an integration on \cal M. In particular, the mass terms of the scalars around ϕα=0\phi_{\alpha}=0 come from the following terms

Tr[12(D¯αϕβ)212ϕαRαβϕβiϕα[F¯αβ,ϕβ]],{\rm Tr}\left[\frac{1}{2}\left(\bar{D}_{\alpha}\phi_{\beta}\right)^{2}-\frac{1}{2}\phi^{\alpha}R_{\alpha\beta}\phi^{\beta}-i\phi^{\alpha}[\bar{F}_{\alpha\beta},\phi^{\beta}]\right], (2.12)

where RαβR_{\alpha\beta} is the Ricci tensor for hαβh_{\alpha\beta} on \cal{M}. Note that we have used the equations of motion for A¯α\bar{A}_{\alpha} in deriving (2.12). On the other hand, the mass terms of the vector fields are provided from the terms,

Tr[12(D¯αaμ+i[ϕα,aμ])2].{\rm Tr}\left[\frac{1}{2}\left(\bar{D}_{\alpha}a_{\mu}+i[\phi_{\alpha},a_{\mu}]\right)^{2}\right]. (2.13)

The second term gives additional contributions to mass at ϕα0\langle\phi_{\alpha}\rangle\neq 0, whose effects we will investigate in sections 4 for =S2{\cal M}=S^{2} and section 5 for =2{\cal M}=\mathbb{CP}^{2}. We show that some of the massive Kaluza-Klein modes become massless by the second term.

2.2 Coset space G/HG/H

In the following, we focus our attention on a compactification on a coset space. See e.g. [40, 25] for more details.

We consider a coset space =G/H{\cal M}=G/H where GG and HH are Lie groups with HGH\subset G. Let us decompose generators of GG as ({ta},{tm})(\{t_{a}\},\{t_{m}\}) where {ta}\{t_{a}\} (a=1,dimHa=1,\cdots\dim H) are a set of generators of HH. Note that {tm}\{t_{m}\} (m=1,,d=dimGdimH)m=1,\cdots,d=\dim G-\dim H) correspond to a basis of the tangent space of G/HG/H. We assume that the generators ta,tmt_{a},t_{m} satisfy the following commutation relations

[ta,tb]=ifctcab,[ta,tm]=ifntnam,[tm,tn]=ifatamn.[t_{a},t_{b}]\ =\ if^{c}{}_{ab}t_{c},\hskip 28.45274pt[t_{a},t_{m}]\ =\ if^{n}{}_{am}t_{n},\hskip 28.45274pt[t_{m},t_{n}]\ =\ if^{a}{}_{mn}t_{a}. (2.14)

A coset space whose generators satisfy the commutation relations of this form is said to be symmetric. In the following, we use a,b,ca,b,c for generators of HH and m,nm,n for generators along G/HG/H. The indices m,nm,n also represent those of coordinates of the tangent space on =G/H{\cal M}=G/H. In this paper, we discuss two examples of symmetric coset spaces, namely S2=SU(2)/U(1)S^{2}=SU(2)/U(1) and 2=SU(3)/U(2)\mathbb{CP}^{2}=SU(3)/U(2). In these cases, tat_{a} are represented in terms of block-diagonal matrices, while tmt_{m} are given in terms of block-off-diagonal matrices, and their commutation relations are apparently of the form (2.14). Non-symmetric coset spaces are discussed in [41].

2.3 Metric and background gauge field on G/HG/H

For a given coset space G/HG/H, there is a “natural” choice for the vielbein eαme^{m}_{\alpha} and the background gauge field A¯α\bar{A}_{\alpha}. Suppose we have a local embedding g:yαG/Hg(y)Gg:y^{\alpha}\in G/H\rightarrow g(y)\in G. Then the Maurer-Cartan 1-form g1dgg^{-1}dg restricted on g(G/H)g(G/H) is written as a sum

g1dg=iea(y)ta+iem(y)tm,g^{-1}dg=ie^{a}(y)t_{a}+ie^{m}(y)t_{m}, (2.15)

where eme^{m} gives the natural choice of the vielbein on G/HG/H while eae^{a} provides the gauge field on the coset space. Indeed, A=eataA=e^{a}t_{a} transforms under gghg\rightarrow gh for hHh\in H, which is a gauge transformation as explained in Appendix C, as Ah1Ahih1dhA\rightarrow h^{-1}Ah-ih^{-1}dh. In the following, we will consider gauge group GYMG_{\rm YM} that includes HH as HGYMH\subset G_{\rm YM} and define the background gauge field on the coset space by

A¯=A¯αdyα:=eαa(y)Tadyα,\bar{A}=\bar{A}_{\alpha}\ dy^{\alpha}:=\ e^{a}_{\alpha}(y)T_{a}dy^{\alpha}, (2.16)

where TaT_{a} are generators of the gauge group GYMG_{\rm YM} which are the corresponding embedding of the generators tat_{a} of HH into the Lie algebra 𝔤YM\mathfrak{g}_{\rm YM} of GYMG_{\rm YM}. In this paper, we consider various different embeddings of HH into GYMG_{\rm YM} for G/H=S2G/H=S^{2} and 2\mathbb{CP}^{2}.

Interestingly, this background gauge field A¯α\bar{A}_{\alpha} automatically satisfies the equations of motion

D¯αF¯αβ=αF¯αβ+i[A¯α,F¯αβ]= 0\bar{D}^{\alpha}\bar{F}_{\alpha\beta}\ =\ \nabla^{\alpha}\bar{F}_{\alpha\beta}+i[\bar{A}^{\alpha},\bar{F}_{\alpha\beta}]\ =\ 0 (2.17)

with respect to the vielbein eαme^{m}_{\alpha} [42]. This can be checked as follows. First, the spin connection ωαnm\omega_{\alpha}{}^{m}{}_{n} defined by dem=ωmnende^{m}=-\omega^{m}{}_{n}\wedge e^{n}, is obtained from the relation

d(g1dg)=g1dgg1dgd(g^{-1}dg)=-g^{-1}dg\wedge g^{-1}dg (2.18)

or equivalently, from the relation (C.11) as

ωα=mnfmeαaan.\omega_{\alpha}{}^{m}{}_{n}\ =\ -f^{m}{}_{an}e^{a}_{\alpha}. (2.19)

Thus, the spin connection is written in terms of the component of the background gauge field eαae^{a}_{\alpha}, and the covariant derivative αF¯αβ\nabla^{\alpha}\bar{F}_{\alpha\beta} with respect to the metric on G/HG/H has the same form as the second term in (2.17). Second, by using the equation (C.10), the field strength of A¯α\bar{A}_{\alpha} turns out to be

F¯αβ=eαmeβnfaTamn.\bar{F}_{\alpha\beta}\ =\ e^{m}_{\alpha}e^{n}_{\beta}f^{a}{}_{mn}T_{a}. (2.20)

Thus, the gauge field strength is non-vanishing in the HH subgroup of GYMG_{\rm YM}. Inserting these expressions of (2.19) and (2.20) into (2.17), we find that it reduces to the Jacobi identity for the structure constants and the background gauge field indeed satisfies the equations of motion. See Appendix D for more details.

2.4 Covariant derivative on G/HG/H

Since the spin connection ωαnm\omega_{\alpha}{}^{m}{}_{n} and the background gauge field A¯α\bar{A}_{\alpha} are given in terms of the same quantity eαae^{a}_{\alpha}, the covariant derivative of ϕm:=emαϕα\phi_{m}:=e^{\alpha}_{m}\phi_{\alpha} can be written as

D¯αϕm:=αϕmωαϕnnm+i[A¯α,ϕm]=αϕm+ieαa(ifnϕnam+[Ta,ϕm]).\bar{D}_{\alpha}\phi_{m}\ :={\partial_{\alpha}\phi_{m}-\omega_{\alpha}{}^{n}{}_{m}\phi_{n}+i[\bar{A}_{\alpha},\phi_{m}]}=\ \partial_{\alpha}\phi_{m}+ie^{a}_{\alpha}\left(-if^{n}{}_{am}\phi_{n}+[T_{a},\phi_{m}]\right). (2.21)

This shows that the field ϕm\phi_{m} can be regarded as a field on the flat d\mathbb{R}^{d} which couples to a gauge field eαae^{a}_{\alpha} as a tensor product of two representations. Actually, the second commutation relation in (2.14) implies that tmt_{m} form a representation RtR_{t} of HH on which the generators are given by ifnamif^{n}{}_{am}. Therefore, ϕm\phi_{m} belongs to the tensor product representation of RtR_{t} and the adjoint representation of GYMG_{\rm YM} and can be decomposed into various irreducible representations of HH. This property plays an important role in the investigations of mass spectrum of various fields with different spins and charges on G/HG/H.

Besides the beautiful properties we have seen above, there are further advantages in choosing a symmetric coset space G/HG/H as the internal manifold \cal M. Most importantly, many properties of the complete set of functions on G/HG/H are well-known and we can explicitly perform the Kaluza-Klein reduction of any field on G/HG/H [40]. For the coset space S2S^{2}, these functions are given by the monopole harmonics [43]. For a general coset space G/HG/H, the Peter-Weyl theorem tells us that each mode functions in the complete set on G/HG/H is labeled by a representation of GG. As mentioned above, the field ϕm\phi_{m} on G/HG/H can be regarded as belonging to a particular representation of HH. This information can be incorporated by taking into account the irreducible decomposition of the representation of GG with respect to HH. See Appendix F for more details.

By using the mode functions, the mass of each Kaluza-Klein mode in the four-dimensional sense can be obtained explicitly [36, 31]. As the mode functions are labeled by the representation of GG and its decomposition with respect to HH, the mass is given in terms of group-theoretic quantities. Namely, it is given in terms the second Casimir invariants of certain representations. We review it in Appendix G, which will be used in the proof of the tachyonic behavior of the symmetric Higgs field observed in the following sections.

3 Symmetric Higgs fields

The Kaluza-Klein reduction of a higher-dimensional Yang-Mills theory on \cal M contains infinitely many fields. If one wants to employ this theory for phenomenological model buildings, it is natural to truncate the theory so that the resulting theory contains only a finite number of light fields. If \cal M is a torus, for example, the lowest mass state is given by a constant mode on \cal M for a scalar field.

When \cal M is a coset space, symmetric fields defined below will provide such lowest mass states [25] on =G/H{\cal M}=G/H. A field ϕm\phi_{m} on G/HG/H is called a symmetric field if its value at yG/Hy^{\prime}\in G/H is related to the value at any other point yG/Hy\in G/H through a local gauge transformation U(y,g0)HGYMU(y,g_{0})\in H\subset G_{\rm YM} and a local Lorentz transformation Λmn(y,g0)\Lambda_{mn}(y,g_{0}) as

ϕm(y)=ΛmnUϕn(y)U,\phi_{m}(y^{\prime})\ =\ \Lambda_{mn}U\phi_{n}(y)U^{\dagger}, (3.1)

where g0g_{0} is an isometry of G/HG/H relating the points y,yy,y^{\prime} [39]. It is a natural generalization of a constant field on a flat space to a coset space. Restricting a higher dimensional theory on symmetric fields on G/HG/H corresponds to focussing on invariant functions under an isometry of G/HG/H up to local symmetry transformations. This criterion is based on the expectation that the lowest energy field configuration is the most symmetric one, and the coset space dimensional reduction retaining only symmetric fields is a natural generalization of the ordinary dimensional reduction retaining only constant modes on a flat torus. Non-constant modes, i.e., non-symmetric fields, correspond to massive fields whose excitation typically costs some amount of energy.

In this paper, instead of restricting the higher dimensional Yang-Mills theory on G/HG/H to only the low lying states, we will keep all higher Kaluza-Klein modes and investigate their important roles in restoration of gauge symmetries, which would be spontaneously broken by condensation of a symmetric field. In particular, we show that some higher excited states become massless under the condensation of a tachyonic symmetric field.

Let us consider a field ϕm\phi_{m} satisfying the condition

D¯αϕm= 0,αϕm= 0.\bar{D}_{\alpha}\phi_{m}\ =\ 0,\hskip 28.45274pt\partial_{\alpha}\phi_{m}\ =\ 0. (3.2)

This turns out to be a symmetric field. See Appendix H for the proof. We call such a field a symmetric Higgs field. The name comes from the fact that the field satisfying the above conditions always has a negative mass squared (G.14), as shown in Appendix G, and develops a vacuum expectation value (vev), which would lead to spontaneous gauge symmetry breaking.

For the symmetric Higgs fields, the scalar potential of S3S_{3} in (LABEL:scalar_potential_original) becomes simplified as

V(ϕ)=14Tr(F¯mn+i[ϕm,ϕn])2,V(\phi)\ =\ \frac{1}{4}{\rm Tr}\left(\bar{F}_{mn}+i[\phi_{m},\phi_{n}]\right)^{2}, (3.3)

where the background field strength (2.20) is

F¯mn=faTamn.\bar{F}_{mn}\ =\ f^{a}{}_{mn}T_{a}. (3.4)

It is nonvanishing only for Ta𝔥𝔤YMT_{a}\in\mathfrak{h}\subset\mathfrak{g}_{\rm YM}, where 𝔥\mathfrak{h} is the Lie algebra of HH.

Recalling the expression for the covariant derivative (2.21), we find that the defining relations (3.2) of symmetric Higgs fields imply

[Ta,ϕm]=ifnϕnam.[T_{a},\phi_{m}]\ =\ if^{n}{}_{am}\phi_{n}. (3.5)

Note that TaT_{a} are generators of GYMG_{\rm YM}, while fnamf^{n}{}_{am} are structure constants of the Lie algebra 𝔤\mathfrak{g} of GG, not those of 𝔤YM\mathfrak{g}_{\rm YM}. Comparing (3.5) with the second equation of (2.14), we can see that ϕm\phi_{m} is expressed by the representation RtR_{t} of HH, possibly with a multiplicity. Thus we can write ϕm\phi_{m} of a symmetric Higgs field as

ϕm(x)=φs(x)Tms,\phi_{m}(x)\ =\ \varphi_{s}(x)T_{m}^{s}, (3.6)

where TmsT^{s}_{m} are generators of 𝔤YM\mathfrak{g}_{\rm YM} satisfying [Ta,Tms]=ifnTnsam.[T_{a},T_{m}^{s}]=if^{n}{}_{am}T_{n}^{s}. Note that TmsT^{s}_{m} are different generators for different ss, as we will see in the following sections. To find the expression for a symmetric Higgs field, we decompose the adjoint representation of GYMG_{\rm YM} into irreducible representations of HH. There could exist representations isomorphic to RtR_{t} in the decomposition. In the following sections, we will explicitly investigate this in various examples.

4 Higgs condensation on S2=SU(2)/U(1)S^{2}={\rm SU}(2)/{\rm U}(1)

In this section, we consider SU(3){\rm SU}(3) Yang-Mills theory compactified on the coset space S2=SU(2)/U(1)S^{2}={\rm SU}(2)/{\rm U}(1). Thus, GYM=SU(3)G_{\rm YM}={\rm SU}(3), G=SU(2)G={\rm SU}(2) and H=U(1)H={\rm U}(1). We choose the generators of su(2)su(2) such that the commutation relations are

[t3,t±]=±t±,[t+,t]= 2t3.[t_{3},t_{\pm}]\ =\ \pm t_{\pm},\hskip 28.45274pt[t_{+},t_{-}]\ =\ 2t_{3}. (4.1)

Then, the index mm for the tangent space takes ++ and -. We denote the generator of u(1)u(1) embedded into su(3)su(3) by TT. The background gauge field is only present in the subgroup H=U(1)H={\rm U}(1), and the scalar potential (3.3) becomes

V(ϕ)\displaystyle V(\phi) =\displaystyle= 18Tr(F¯++i[ϕ+,ϕ])2\displaystyle-\frac{1}{8}{\rm Tr}\left(\bar{F}_{+-}+i[\phi_{+},\phi_{-}]\right)^{2} (4.2)
=\displaystyle= 18Tr(2T[ϕ+,ϕ])2.\displaystyle\frac{1}{8}{\rm Tr}\left(2T-[\phi_{+},\phi_{-}]\right)^{2}.

In the following, we will show that different choices of TT give us different contents of symmetric Higgs fields with different patterns of their condensation.

When the coset space is S2S^{2}, the background gauge field A¯α\bar{A}_{\alpha} and the zweibein eαme^{m}_{\alpha} can be explicitly written as reviewed in Appendix E. In fact, A¯α\bar{A}_{\alpha} is given by the monopole configuration on S2S^{2} embedded into SU(3){\rm SU}(3) gauge group. Details on these expressions, in addition to the explicit formula for the covariant derivative D¯mϕn\bar{D}_{m}\phi_{n}, can be found in Appendix E. We can use these explicit expressions, in particular, the monopole harmonics [43] to investigate the spectrum in the Kaluza-Klein reduction. However it will turn out that a more abstract formalism [40] reviewed in Appendix F is sufficient for the purpose since various analytic calculations can be reduced to group-theoretic arguments on the coset space. Such an abstract formalism is straightforwardly extended to more general coset spaces, such as 2=SU(3)/(SU(2)×U(1))\mathbb{CP}^{2}={\rm SU}(3)/({\rm SU}(2)\times{\rm U(1)}) which will be discussed in the next section.

4.1 Embedding of H=H=\,U(1)(1) into GYM=G_{\rm YM}=\,SU(3)(3): Case 1

Our first choice of the embedding of the H=U(1)H=\rm U(1) generator TT in GYM=SU(3)G_{\rm YM}=\rm SU(3) is

T=12[100010000].T\ =\ \frac{1}{2}\left[\begin{array}[]{ccc}1&0&0\\ 0&-1&0\\ 0&0&0\end{array}\right]. (4.3)

The background flux F¯+=2iT\bar{F}_{+-}=-2iT breaks the gauge group SU(3){\rm SU}(3) to its Cartan subgroup U(1)×U(1){\rm U}(1)\times{\rm U}(1).

Let us now find the symmetric Higgs field satisfying the relation (3.5) for the U(1) generator TT. We first define TT-charges qijq_{ij} of ϕ±,ij\phi_{\pm,ij} fields by

[T,ϕ±]ij=qijϕ±,ij,[T,\phi_{\pm}]_{ij}\ =\ q_{ij}\,\phi_{\pm,ij}, (4.4)

where i,j=1,2,3i,j=1,2,3 are indices of 3×33\times 3 matrices, and no summation is taken. For the choice (4.3) of TT, the TT-charges are given in the matrix notation as

q=[0112101212120].q\ =\ \left[\begin{array}[]{ccc}0&1&\frac{1}{2}\\ -1&0&-\frac{1}{2}\\ -\frac{1}{2}&\frac{1}{2}&0\end{array}\right]. (4.5)

Then, the condition (3.5) for the symmetric Higgs field and the commutation relation (4.1) tell us that the (i,j)=(1,2)(i,j)=(1,2) and (2,1)(2,1) components of ϕ±,ij\phi_{\pm,ij} with TT-charge ±1\pm 1 provide us with the symmetric Higgs fields. Thus there is only one symmetric Higgs field (and its complex conjugate) given by

ϕ+(x)=[0φ(x)0000000],ϕ(x)=[000φ(x)00000],\phi_{+}(x)\ =\ \left[\begin{array}[]{ccc}0&\varphi(x)&0\\ 0&0&0\\ 0&0&0\end{array}\right],\hskip 28.45274pt\phi_{-}(x)\ =\ \left[\begin{array}[]{ccc}0&0&0\\ \varphi^{\dagger}(x)&0&0\\ 0&0&0\end{array}\right], (4.6)

where we have used ϕ=(ϕ+)\phi_{-}=(\phi_{+})^{\dagger}.

Inserting these expressions into the scalar potential (4.2), we obtain the scalar potential for the symmetric Higgs field φ\varphi

V(ϕ)=14(1|φ|2)2.V(\phi)\ =\ \frac{1}{4}\left(1-|\varphi|^{2}\right)^{2}. (4.7)

Thus φ\varphi will acquire vev at |φ|=1|\varphi|=1. At the origin φ=0\varphi=0, as mentioned before, the gauge symmetry SU(3)\rm SU(3) is broken to U(1)×U(1)SU(3){\rm U}(1)\times{\rm U}(1)\subset{\rm SU}(3) by the background gauge flux. When the symmetric Higgs field acquires vev at |φ|=1|\varphi|=1, the gauge symmetry is expected to be further broken to U(1){\rm U}(1) by the Higgs mechanism. Thus the expected symmetry breaking pattern is as follows:

GYM=SU(3)backgroundfluxU(1)×U(1)HiggsvevU(1)?G_{\rm YM}={\rm SU}(3)\xrightarrow{{\rm background}\ {\rm flux}}{\rm U}(1)\times{\rm U}(1)\xrightarrow{{\rm Higgs}\ {\rm vev}}{\rm U}(1)\ ? (4.8)

This is the usual argument for the gauge symmetry breaking in the context of the coset space dimensional reduction in which only the low lying states are taken into considerations. However, the conclusion of the gauge symmetry breaking is suspicious in view of the higher dimensional gauge theory with the Kaluza-Klein reduction. The reason is the following. Note that we have vanishing scalar potential V(|φ|=1)=0V({|\varphi|=1})=0 at the the global minimum of V(ϕ)V(\phi). Since the scalar potential originally comes from the terms (LABEL:scalar_potential_original), the vanishing scalar potential implies that the gauge field AαA_{\alpha} at the symmetric Higgs vev |φ|=1|\varphi|=1 must be a pure gauge, and we must conclude that the full gauge symmetry SU(3){\rm SU}(3) is recovered at the symmetric Higgs vacuum, instead of being broken to U(1).

In the rest of this section, in order to show the restoration of the gauge symmetry, we will explicitly see that some of the originally massive Kaluza-Klein vector fields become massless at vev |φ|=1|\varphi|=1, and eight massless vector fields emerge at the symmetric Higgs vacuum. These massless vector fields are the gauge fields due to the general argument by Weinberg [44].

Mass term of the vector field aμa_{\mu} comes from the term (2.13), and a vector field is massless in the presence of the symmetric Higgs vev if and only if

D¯+aμ(x,y)+i[T+,aμ(x,y)]= 0,T+:=[010000000]\bar{D}_{+}a_{\mu}(x,y)+i[T_{+},a_{\mu}(x,y)]\ =\ 0,\hskip 28.45274ptT_{+}\ :=\ \left[\begin{array}[]{ccc}0&1&0\\ 0&0&0\\ 0&0&0\end{array}\right] (4.9)

is satisfied. Note that TT, T+T_{+} and T:=(T+)T_{-}:=(T_{+})^{\dagger} form an su(2)su(2) subalgebra of su(3)su(3), and aμa_{\mu} is in the adjoint representation 𝟖\bf 8 of su(3)su(3). Thus, by the irreducible decomposition of 𝟖\bf 8 of su(3)su(3) into 𝟑𝟐𝟐𝟏\bf 3\oplus 2\oplus 2^{\prime}\oplus 1 of su(2)su(2), the condition (4.9) can be decomposed into the following four conditions. First, for the representation 𝟑\bf 3, we have

[1]MasslessCond.for 3D¯+[aμ,12aμ,11aμ,22aμ,21]=i[aμ,22aμ,112aμ,210].[1]\ {\rm Massless\ Cond.\ for\ {\bf 3}}\hskip 28.45274pt\bar{D}_{+}\left[\begin{array}[]{c}a_{\mu,12}\\[2.84526pt] a_{\mu,11}-a_{\mu,22}\\[2.84526pt] a_{\mu,21}\end{array}\right]\ =\ -i\left[\begin{array}[]{c}a_{\mu,22}-a_{\mu,11}\\[2.84526pt] 2a_{\mu,21}\\[2.84526pt] 0\end{array}\right]. (4.10)

For the representations 𝟐\bf 2 and 𝟐\bf 2^{\prime}, we have

[2]MasslessCond.for 2D¯+[aμ,13aμ,23]=i[aμ,230][2]\ {\rm Massless\ Cond.\ for}\ {\bf 2}\hskip 28.45274pt\bar{D}_{+}\left[\begin{array}[]{c}a_{\mu,13}\\[2.84526pt] a_{\mu,23}\end{array}\right]\ =\ -i\left[\begin{array}[]{c}a_{\mu,23}\\[2.84526pt] 0\end{array}\right] (4.11)

with the condition for their conjugate components aμ,31,aμ,32a_{\mu,31},a_{\mu,32} which is equivalent to

[2]MasslessCond.for 2D¯[aμ,13aμ,23]=i[0aμ,13],[2^{\prime}]\ {\rm Massless\ Cond.\ for}\ {\bf 2^{\prime}}\hskip 28.45274pt\bar{D}_{-}\left[\begin{array}[]{c}a_{\mu,13}\\[2.84526pt] a_{\mu,23}\end{array}\right]\ =\ -i\left[\begin{array}[]{c}0\\[2.84526pt] a_{\mu,13}\end{array}\right], (4.12)

and finally,

[3]MasslessCond.for 1D¯+aμ,33= 0[3]\ {\rm Massless\ Cond.\ for\ {\bf 1}}\hskip 28.45274pt\bar{D}_{+}a_{\mu,33}\ =\ 0 (4.13)

for the singlet representation 𝟏\bf 1. A vector field satisfying one of these conditions become massless at the global minimum of the Higgs potential V(φ)V(\varphi) at |φ|=1|\varphi|=1. These are a set of first-order partial differential equations which can be written explicitly by using the formulas in Appendix E, and the number of massless vector fields can be found by solving the above differential equations. In the following, instead of solving them explicitly, we solve these conditions by reducing to a group-theoretic problem.

For this purpose, we need to understand the action of the covariant derivative D¯±\bar{D}_{\pm} on aμa_{\mu} [40]. The action can be simplified by choosing a suitable complete set of functions on S2S^{2} which can be used to expand aμ(x,y)a_{\mu}(x,y). Generally speaking, as explained in Appendix F, due to the Peter-Weyl theorem, a complete set of functions on a group manifold GG is given by the representation matrices ρR(g)IJ\rho^{R}(g)_{IJ} for all the representation RR of GG and their components I,J=1,,dimRI,J=1,\cdots,\dim R. Then a complete set of functions on G/HG/H is obtained by imposing particular transformation laws under HH, corresponding to the TT-charge of functions on G/HG/H. Collecting all the representations of HH, the complete set on GG is recovered.

In the case of S2=SU(2)/U(1)S^{2}={\rm SU}(2)/{\rm U}(1), a complete set on S2S^{2}, collecting all the charges of H=U(1)H={\rm U}(1), is given by

fmmj(y)wherej= 0,12, 1,32,,jm,mj.f^{j}_{mm^{\prime}}(y)\hskip 14.22636pt{\rm where}\ \ j\ =\ 0,\ \frac{1}{2},\ 1,\ \frac{3}{2},\ \cdots,\hskip 14.22636pt-j\ \leq\ m,m^{\prime}\,\leq\ j. (4.14)

Each jj corresponds to the spin jj representation of SU(2){\rm SU}(2). As explained in Appendix F, the function fmmjf^{j}_{mm^{\prime}} has TT-charge mm of H=U(1)H={\rm U}(1). Thus, m=0m=0 gives the usual spherical harmonics, while m0m\neq 0 modes are the monopole spherical harmonics with TT-charge mm, which are relevant in the monopole background.

A field χ(y)\chi(y) on S2S^{2} with the TT-charge qq is then expanded in terms of fqmj(y)f^{j}_{qm^{\prime}}(y) as

χ(y)=jm=jjcmjfqmj(y),\chi(y)\ =\ \sum_{j}\sum_{m^{\prime}=-j}^{j}c^{j}_{m^{\prime}}f^{j}_{qm^{\prime}}(y), (4.15)

where the sum of jj is taken over all values of the spin jj whose magnetic quantum number mm^{\prime} can take qq. Explicitly, jj in the sum must satisfy

jqj,jq.-j\ \leq\ q\ \leq\ j,\hskip 28.45274ptj-q\ \in\ \mathbb{Z}. (4.16)

From this expansion, we obtain 2j+12j+1 complex-valued fields, labeled by mm^{\prime}, with the TT-charge qq from each jj.

In order to discuss the massless condition (4.9) for vector fields, it is sufficient to know the action of D¯+\bar{D}_{+} on the mode functions fmmj(y)f^{j}_{mm^{\prime}}(y). From (G.3) in Appendix G, this action turns out to be given by

D¯+fmmj(y)=in=jj(T+(j))mnfnmj(y),\bar{D}_{+}f^{j}_{mm^{\prime}}(y)\ =\ -i\sum_{n=-j}^{j}\left(T^{(j)}_{+}\right)_{mn}f^{j}_{nm^{\prime}}(y), (4.17)

where T+(j)T^{(j)}_{+} is the spin-jj representation of t+t_{+}. Note that it is valid irrespective of the value of the symmetric Higgs field. Thus, the condition (4.9) for massless vector fields is reduced to algebraic relations of the coefficients cμ,i1i2,mjc^{j}_{\mu,i_{1}i_{2},m^{\prime}} in the mode expansion

aμ,i1i2(y)=jm=jjcμ,i1i2,mjfq(i1,i2),mj(y)a_{\mu,i_{1}i_{2}}(y)\ =\ \sum_{j}\sum_{m^{\prime}=-j}^{j}c^{j}_{\mu,i_{1}i_{2},m^{\prime}}f^{j}_{q(i_{1},i_{2}),m^{\prime}}(y) (4.18)

between the first and the second terms in (4.9). Here q(i1,i2)q(i_{1},i_{2}) is the TT-charge of (i1,i2)(i_{1},i_{2})-component of aμa_{\mu}. The first term in (4.9) is a multiplication of T+(j)T^{(j)}_{+} on the complete set fmmj(y)f^{j}_{mm^{\prime}}(y) due to the covariant derivative D¯+\bar{D}_{+} while the second term is the adjoint action of T+T_{+} due to the symmetric Higgs vev. If we can choose the expansion coefficients of aμa_{\mu} such that these two actions have the same effect, then we obtain a massless vector field.

Let us check whether this condition can be satisfied. First, we consider (4.11) of the massless condition for representation 𝟐\bf 2, i.e., q=±1/2q=\pm 1/2. Thus, the representations of SU(2)\rm SU(2) are restricted to be j=k+1/2j=k+1/2 for non-negative integers kk. This can be written as

k=0m=k12k+12cmk(T+(k+12))12,nfnmk+12(y)\displaystyle\sum_{k=0}^{\infty}\sum_{m^{\prime}=-k-\frac{1}{2}}^{k+\frac{1}{2}}c^{k}_{m^{\prime}}\left(T^{(k+\frac{1}{2})}_{+}\right)_{\frac{1}{2},n}f^{k+\frac{1}{2}}_{nm^{\prime}}(y) =\displaystyle= k=0m=k12k+12c~mkf12,mk+12(y)\displaystyle\sum_{k=0}^{\infty}\sum_{m^{\prime}=-k-\frac{1}{2}}^{k+\frac{1}{2}}\tilde{c}^{k}_{m^{\prime}}f^{k+\frac{1}{2}}_{-\frac{1}{2},m^{\prime}}(y) (4.19)

and

k=0m=k12k+12c~mk(T+(k+12))12,nfnmk+12(y)\displaystyle\sum_{k=0}^{\infty}\sum_{m^{\prime}=-k-\frac{1}{2}}^{k+\frac{1}{2}}\tilde{c}^{k}_{m^{\prime}}\left(T^{(k+\frac{1}{2})}_{+}\right)_{-\frac{1}{2},n}f^{k+\frac{1}{2}}_{nm^{\prime}}(y) =\displaystyle= 0,\displaystyle 0, (4.20)

where we renamed the coefficients cμ,13,mjc^{j}_{\mu,13,m^{\prime}} and cμ,23,mjc^{j}_{\mu,23,m^{\prime}} as cmkc^{k}_{m^{\prime}} and c~mk\tilde{c}^{k}_{m^{\prime}}, respectively. These equations are satisfied if and only if cm0=c~m0c^{0}_{m^{\prime}}=\tilde{c}^{0}_{m^{\prime}} are the only non-zero coefficients. Note that in our normalization and notation

(T+(12))12,12=(σ+)12= 1,\left(T^{(\frac{1}{2})}_{+}\right)_{\frac{1}{2},-\frac{1}{2}}\ =(\sigma_{+})_{12}=\ 1\,, (4.21)

where σ+\sigma_{+} is the Pauli matrix. The same coefficients also solve (4.12). Consequently, we have shown that

aμ,13(x,y)=m=±12cm0(x)f12,m12(y),aμ,23(x,y)=m=±12cm0(x)f12,m12(y)a_{\mu,13}(x,y)\ =\ \sum_{m^{\prime}=\pm\frac{1}{2}}c^{0}_{m^{\prime}}(x)f^{\frac{1}{2}}_{\frac{1}{2},m^{\prime}}(y),\hskip 28.45274pta_{\mu,23}(x,y)\ =\ \sum_{m^{\prime}=\pm\frac{1}{2}}c^{0}_{m^{\prime}}(x)f^{\frac{1}{2}}_{-\frac{1}{2},m^{\prime}}(y) (4.22)

and their conjugates give us four massless vector fields.

Next, we consider (4.10) for the massless condition for representation 𝟑\bf 3, i.e., q=0,±1q=0,\pm 1. Thus, the representations of SU(2)\rm SU(2) are restricted to be j=lj=l for non-negative integers ll. The range of ll depends on the TT-charges. The linear combinations

aμ,12,12(aμ,11aμ,22),aμ,21-a_{\mu,12},\hskip 28.45274pt\frac{1}{\sqrt{2}}(a_{\mu,11}-a_{\mu,22}),\hskip 28.45274pta_{\mu,21} (4.23)

as the independent fields are convenient for our purpose. Then, the condition (4.10) can be written as

l=1m=llcml(T+(l))1,nfnml(y)\displaystyle\sum_{l=1}^{\infty}\sum_{m^{\prime}=-l}^{l}c^{l}_{m^{\prime}}\left(T^{(l)}_{+}\right)_{1,n}f^{l}_{nm^{\prime}}(y) =\displaystyle= 2l=0m=llc~mlf0,ml(y),\displaystyle\sqrt{2}\sum_{l=0}^{\infty}\sum_{m^{\prime}=-l}^{l}\widetilde{c}^{l}_{m^{\prime}}f^{l}_{0,m^{\prime}}(y), (4.24)
l=0m=llc~ml(T+(l))0,nfnml(y)\displaystyle\sum_{l=0}^{\infty}\sum_{m^{\prime}=-l}^{l}\widetilde{c}^{l}_{m^{\prime}}\left(T^{(l)}_{+}\right)_{0,n}f^{l}_{nm^{\prime}}(y) =\displaystyle= 2l=1m=llc^mlf1,ml(y),\displaystyle\sqrt{2}\sum_{l=1}^{\infty}\sum_{m^{\prime}=-l}^{l}\widehat{c}^{l}_{m^{\prime}}f^{l}_{-1,m^{\prime}}(y), (4.25)
l=1m=llc^ml(T+(l))1,nfnml(y)\displaystyle\sum_{l=1}^{\infty}\sum_{m^{\prime}=-l}^{l}\widehat{c}^{l}_{m^{\prime}}\left(T^{(l)}_{+}\right)_{-1,n}f^{l}_{nm^{\prime}}(y) =\displaystyle= 0.\displaystyle 0. (4.26)

These equations are satisfied if and only if cm1=c~m1=c^m1c^{1}_{m^{\prime}}=\widetilde{c}^{1}_{m^{\prime}}=\widehat{c}^{1}_{m^{\prime}} are the only non-zero coefficients. Then, the following three combinations

aμ,12\displaystyle-a_{\mu,12} =\displaystyle= m=1,0,1cm1f1,m1(y),\displaystyle{\sum_{m^{\prime}=-1,0,1}c^{1}_{m^{\prime}}f^{1}_{1,m^{\prime}}(y),} (4.27)
12(aμ,11aμ,22)\displaystyle\frac{1}{\sqrt{2}}(a_{\mu,11}-a_{\mu,22}) =\displaystyle= m=1,0,1cm1f0,m1(y),\displaystyle\sum_{m^{\prime}=-1,0,1}c^{1}_{m^{\prime}}f^{1}_{0,m^{\prime}}(y), (4.28)
aμ,21\displaystyle a_{\mu,21} =\displaystyle= m=1,0,1cm1f1,m1(y),\displaystyle\sum_{m^{\prime}=-1,0,1}c^{1}_{m^{\prime}}f^{1}_{-1,m^{\prime}}(y), (4.29)

with the condition aμ=aμa_{\mu}^{\dagger}=a_{\mu}, give us 3 massless vector fields. In fact, this can be easily anticipated by rewriting (4.10) as

D¯+[aμ,1212(aμ,11aμ,22)aμ,21]=iT+(1)[aμ,1212(aμ,11aμ,22)aμ,21].\bar{D}_{+}\left[\begin{array}[]{c}-a_{\mu,12}\\[2.84526pt] \frac{1}{\sqrt{2}}(a_{\mu,11}-a_{\mu,22})\\[2.84526pt] a_{\mu,21}\end{array}\right]\ =\ -iT^{(1)}_{+}\left[\begin{array}[]{c}-a_{\mu,12}\\[2.84526pt] \frac{1}{\sqrt{2}}(a_{\mu,11}-a_{\mu,22})\\[2.84526pt] a_{\mu,21}\end{array}\right]. (4.30)

Namely, these three components form the triplet of the su(2)su(2) subalgebra, as mentioned before. Note that aμ,11aμ,22a_{\mu,11}-a_{\mu,22} has also a contribution from j=0j=0 which was massless before the Higgs condensation. This becomes massive due to the Higgs mechanism.

Finally, let us consider the condition (4.13) for the massless condition for representation 𝟏\bf 1, i.e., q=0q=0. Thus, the representation fm,mjf^{j}_{m,m^{\prime}} of SU(2)SU(2) is restricted to be j=lj=l for non-negative integers ll. The condition simply means that aμ,33a_{\mu,33} is independent of yy, resulting in one massless vector field. This is nothing but the U(1){\rm U}(1) gauge field which is unbroken after the Higgs condensation.

In total, we have found eight massless vector fields which should correspond to the SU(3){\rm SU}(3) gauge field which is expected to appear at the symmetric Higgs vacuum. Therefore we conclude that, contrary to the expectation in (4.8) within the analysis of the low lying states, the symmetry breaking-restoration pattern is given by

GYM=SU(3)backgroundfluxU(1)×U(1)HiggsvevSU(3)G_{\rm YM}={\rm SU}(3)\xrightarrow{{\rm background}\ {\rm flux}}{\rm U}(1)\times{\rm U}(1)\xrightarrow{{\rm Higgs}\ {\rm vev}}{\rm SU}(3)\ (4.31)

if all the Kaluza-Klein modes are taken into considerations. It should be noted that seven massless vector fields out of eight ones come from the Kaluza-Klein modes which were massive before the Higgs condensation. Indeed, all the massless vector fields before the Higgs condensation come from j=0j=0 mode in the expansions, while seven massless fields after the Higgs condensation come from the j=12j=\frac{1}{2} and j=1j=1 modes. This phenomenon happens because we keep all the Kaluza-Klein modes in the model, in contrast to the simplest analysis of the coset space dimensional reduction in which only the low lying modes are taken into account. Usually, the Kaluza-Klein modes are considered to be so heavy that they are not considered in discussing the dynamics of light fields. However, since our model has only a single mass scale given by the radius of the coset space, the potential height and the Higgs vev are also of the same Kaluza-Klein mass scale. This enables some of the massive Kaluza-Klein modes to become massless.

The investigation of the massless vector fields performed above is possible because the vev of φ\varphi is exactly the right value so that the vev of ϕ+\phi_{+} is equal to the generator T+T_{+}, including the overall normalization. We will observe in the following that this coincidence persists for the other models discussed in this paper. It is very interesting to clarify whether this is the general feature of the Kaluza-Klein reduction on coset spaces. If this is the case, the masses of the fields at the symmetric Higgs vacuum would be possibly given in terms of group-theoretic quantities, as the mass formulas in Appendix G are valid before the symmetric Higgs condensation.

4.2 Embedding of H=H=\,U(1)(1) into GYM=G_{\rm YM}=\,SU(3)(3): Case 2

Our next choice of an embedding of U(1) charge TT into GYM=G_{\rm YM}=SU(3) is

T=13[112].T\ =\ \frac{1}{3}\left[\begin{array}[]{ccc}1&&\\ &1&\\ &&-2\end{array}\right]. (4.32)

The corresponding background flux F¯+=2iT\bar{F}_{+-}=-2iT breaks the gauge group GYM=SU(3)G_{\rm YM}={\rm SU}(3) into SU(2)×U(1){\rm SU}(2)\times{\rm U}(1).

The TT-charges for the components of ϕ+\phi_{+} defined in (4.4) are given by

q=[001001110].q\ =\ \left[\begin{array}[]{ccc}0&0&1\\ 0&0&1\\ -1&-1&0\end{array}\right]. (4.33)

Recall that the components of ϕ+\phi_{+} with the TT-charge +1+1 become the symmetric Higgs fields, and there are two such components. Therefore, the symmetric Higgs fields are given by

ϕ+=[00φ100φ2000],ϕ=[000000φ1φ20].\phi_{+}\ =\ \left[\begin{array}[]{ccc}0&0&\varphi_{1}\\ 0&0&\varphi_{2}\\ 0&0&0\end{array}\right],\hskip 28.45274pt\phi_{-}\ =\ \left[\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ \varphi_{1}^{\dagger}&\varphi_{2}^{\dagger}&0\end{array}\right]. (4.34)

The components φ1\varphi_{1} and φ2\varphi_{2} form a doublet of the unbroken SU(2){\rm SU}(2) gauge group.

The scalar potential for this symmetric Higgs doublet can be obtained by inserting the above expressions into (4.2). We obtain

V(ϕ)=14(1|φ|2)2+112,V(\phi)\ =\ \frac{1}{4}\left(1-|\varphi|^{2}\right)^{2}+\frac{1}{12}, (4.35)

where |φ|2:=|φ1|2+|φ2|2|\varphi|^{2}:=|\varphi_{1}|^{2}+|\varphi_{2}|^{2}. At the minimum of the potential, they acquire the vev

φ1= 1,φ2= 0,\varphi_{1}\ =\ 1,\hskip 28.45274pt\varphi_{2}\ =\ 0, (4.36)

up to a global SU(2) gauge transformation. This would break the gauge group SU(2)×U(1){\rm SU}(2)\times{\rm U}(1) preserved by the background flux to U(1){\rm U}(1). Manton [8] applied this mechanism of the gauge symmetry breaking to realization of the Weinberg-Salam model based on the six-dimensional Yang-Mills theory. The expected symmetry breaking pattern within the low lying states would be as follows:

GYM=SU(3)backgroundfluxSU(2)×U(1)HiggsvevU(1)?G_{\rm YM}={\rm SU}(3)\xrightarrow{{\rm background}\ {\rm flux}}{\rm SU}(2)\times{\rm U}(1)\xrightarrow{{\rm Higgs}\ {\rm vev}}{\rm U}(1)\ ? (4.37)

In order to perform the calculation of V(ϕ)V(\phi) while keeping the SU(2)×U(1){\rm SU}(2)\times{\rm U}(1) gauge invariance, it is convenient to introduce 3×33\times 3 matrices 𝒯s{\cal T}_{s} (s=1,2s=1,2) and write ϕ+\phi_{+} in (4.34) as

ϕ+=s=1,2φs𝒯s.\phi_{+}\ =\sum_{s=1,2}\varphi_{s}{\cal T}_{s}.\hskip 28.45274pt (4.38)

The matrices 𝒯s{\cal T}_{s} defined by this relation satisfy

[𝒯s,𝒯t]=σts[010000000]+σts+[000100000]+σts3[12000120000]+32δtsT,[{\cal T}_{s},{\cal T}_{t}^{\dagger}]\ =\ \sigma^{-}_{ts}\left[\begin{array}[]{ccc}0&1&0\\ 0&0&0\\ 0&0&0\end{array}\right]+\sigma^{+}_{ts}\left[\begin{array}[]{ccc}0&0&0\\ 1&0&0\\ 0&0&0\end{array}\right]+\sigma^{3}_{ts}\left[\begin{array}[]{ccc}\frac{1}{2}&0&0\\ 0&-\frac{1}{2}&0\\ 0&0&0\end{array}\right]+\frac{3}{2}\delta_{ts}T, (4.39)

where σts+\sigma^{+}_{ts} etc. are (t,s)(t,s)-components of the Pauli matrices.

Let us count the number of massless vector fields at the symmetric Higgs vacuum. We notice that the vev of ϕ+\phi_{+} is

ϕ+=[001000000]\phi_{+}\ =\ \left[\begin{array}[]{ccc}0&0&1\\ 0&0&0\\ 0&0&0\end{array}\right] (4.40)

which is the spin-12\frac{1}{2} representation of t+t_{+} embedded into an su(2)su(2) subalgebra of su(3)su(3) different from the one in the previous section. The massless conditions in this case can be obtained from the previous ones by simply exchanging 22 and 33 in the matrix indices. For example, we have

D¯[aμ,12aμ,32]=i[0aμ,12].\bar{D}_{-}\left[\begin{array}[]{c}a_{\mu,12}\\[2.84526pt] a_{\mu,32}\end{array}\right]\ =\ -i\left[\begin{array}[]{c}0\\[2.84526pt] a_{\mu,12}\end{array}\right]. (4.41)

In this case, however, the TT-charges of the components are different. The TT-charges for aμa_{\mu} are

q=[001001110].q\ =\ \left[\begin{array}[]{ccc}0&0&1\\ 0&0&1\\ -1&-1&0\end{array}\right]. (4.42)

Therefore, aμ,12a_{\mu,12} has the TT-charge 0, while aμ,32a_{\mu,32} has the TT-charge 1-1. Their mode expansions are given as

aμ,12\displaystyle a_{\mu,12} =\displaystyle= j=0m=jjcmjf0,mj,\displaystyle\sum_{j=0}^{\infty}\sum_{m^{\prime}=-j}^{j}c^{j}_{m^{\prime}}f^{j}_{0,m^{\prime}}, (4.43)
aμ,32\displaystyle a_{\mu,32} =\displaystyle= j=1m=jjc~mjf1,mj.\displaystyle\sum_{j=1}^{\infty}\sum_{m^{\prime}=-j}^{j}\tilde{c}^{j}_{m^{\prime}}f^{j}_{-1,m^{\prime}}. (4.44)

We find that D¯f0,mj\bar{D}_{-}f^{j}_{0,m^{\prime}} vanishes only if j=0j=0. Since D¯aμ,32\bar{D}_{-}a_{\mu,32} does not have a contribution from the spin-0 representation, we conclude that the condition (4.41) does not have a solution.

We also have

D¯+[aμ,1312(aμ,11aμ,33)aμ,31]=iT+(1)[aμ,1312(aμ,11aμ,33)aμ,31].\bar{D}_{+}\left[\begin{array}[]{c}-a_{\mu,13}\\[2.84526pt] \frac{1}{\sqrt{2}}(a_{\mu,11}-a_{\mu,33})\\[2.84526pt] a_{\mu,31}\end{array}\right]\ =\ -iT^{(1)}_{+}\left[\begin{array}[]{c}-a_{\mu,13}\\[2.84526pt] \frac{1}{\sqrt{2}}(a_{\mu,11}-a_{\mu,33})\\[2.84526pt] a_{\mu,31}\end{array}\right]. (4.45)

The TT-charge assignment for aμa_{\mu} turns out to be appropriate so that we can find the following solution

aμ,13(x,y)\displaystyle-a_{\mu,13}(x,y) =\displaystyle= m=11cm1f1,m1(y),\displaystyle{\sum_{m^{\prime}=-1}^{1}c^{1}_{m^{\prime}}f^{1}_{1,m^{\prime}}(y),} (4.46)
12(aμ,11(x,y)aμ,33(x,y))\displaystyle\frac{1}{\sqrt{2}}(a_{\mu,11}(x,y)-a_{\mu,33}(x,y)) =\displaystyle= m=11cm1f0,m1(y),\displaystyle{\sum_{m^{\prime}=-1}^{1}c^{1}_{m^{\prime}}f^{1}_{0,m^{\prime}}(y),} (4.47)
aμ,13(x,y)\displaystyle a_{\mu,13}(x,y) =\displaystyle= m=11cm1f1,m1(y).\displaystyle{\sum_{m^{\prime}=-1}^{1}c^{1}_{m^{\prime}}f^{1}_{-1,m^{\prime}}(y).} (4.48)

They give us three massless vector fields.

The last condition

D¯+aμ,22= 0\bar{D}_{+}a_{\mu,22}\ =\ 0 (4.49)

gives us one massless vector field.

In total, we have found four massless vector fields. Three of them were massive before the Higgs condensation. Since scalar potential (4.35) does not vanish at the symmetric Higgs vacuum |φ|=1|\varphi|=1, there remains a non-trivial flux after the Higgs condensation which prevents the full SU(3){\rm SU}(3) gauge symmetry from recovering. Probably, the gauge group at the symmetric Higgs vacuum would be SU(2)×U(1){\rm SU}(2)\times{\rm U}(1), where the SU(2){\rm SU}(2) part is not the one preserved by F¯+\bar{F}_{+-} but an “emergent” one. Therefore we may conclude that, contrary to the expectation in (4.37) within the analysis of the low lying states, the symmetry breaking-restoration pattern is given by

GYM=SU(3)backgroundfluxSU(2)×U(1)HiggsvevSU(2)×U(1).G_{\rm YM}={\rm SU}(3)\xrightarrow{{\rm background}\ {\rm flux}}{\rm SU}(2)\times{\rm U}(1)\xrightarrow{{\rm Higgs}\ {\rm vev}}{\rm SU}(2)\times{\rm U}(1)\,. (4.50)

For confirmation of this pattern, a more detailed analysis will be necessary.

It is natural to ask whether the symmetric Higgs vacuum |φ|=1|\varphi|=1 is stable or not. In the previous section, the stability is obvious since the vacuum attains the global minimum of the scalar potential in every direction of the field space. For the case in this section, it is possible that there still exists a Higgs field at the symmetric Higgs vacuum, and a further condensation would occur.

At least, we can show that |φ|=1|\varphi|=1 is a classical solution of the full theory including all Kaluza-Klein modes. In other words, we claim that the symmetric Higgs vacuum discussed in this section has the same relevance as the trivial solution before the symmetric Higgs condensation which has been discussed in the literature [31]. To show this, we need to confirm that the symmetric Higgs vev does not act as a source for other scalar fields coming from the Kaluza-Klein expansion of ϕ±\phi_{\pm}. If there would exist terms in the scalar potential of the form

Tr(F¯φΦ),Tr(φ3Φ),{\rm Tr}(\bar{F}\varphi\Phi),\hskip 28.45274pt{\rm Tr}(\varphi^{3}\Phi), (4.51)

where Φ\Phi indicates scalar fields other that the symmetric Higgs fields, then the vev of φ\varphi would give a source term of Φ\Phi, so that Φ=0\Phi=0 is not the classical solution. As mentioned above, there could exist terms with Φ2\Phi^{2} which would indicate the presence of other Higgs fields. Since this allows Φ=0\Phi=0 to be a classical solution, we ignore them in the following.

Recall that a symmetric Higgs field φ\varphi is a constant mode on S2S^{2} and is singlet for the HH transformation, which are implied by the conditions (3.2). Then, Φ\Phi is a non-constant mode on S2S^{2} or HH-non-singlet. This implies that the terms of the second kind in (4.51) is absent. Indeed, if Φ\Phi is a non-constant mode, then the integration of Tr(φ3Φ){\rm Tr}(\varphi^{3}\Phi) over S2S^{2}, performed in the Kaluza-Klein reduction, vanishes due to the orthogonality condition for the mode functions fmmjf^{j}_{mm^{\prime}}. On the other hand, if Φ\Phi is HH-non-singlet, then Tr(φ3Φ){\rm Tr}(\varphi^{3}\Phi) simply vanishes since the scalar potential is HH-singlet. The terms of the first kind in (4.51) are also prohibited since the background flux F¯+\bar{F}_{+-} is also constant on S2S^{2} and HH-singlet. The latter is valid since F¯+\bar{F}_{+-} is invariant under the U(1){\rm U}(1) gauge transformation, and is also invariant under the local Lorentz transformation.

4.3 Embedding of H=H=\,U(1)(1) into GYM=G_{\rm YM}=\,SU(3)(3): Case 3

Our last choice for TT is

T=[101].T\ =\ \left[\begin{array}[]{ccc}1&&\\ &0&\\ &&-1\end{array}\right]. (4.52)

The background flux F¯+=2iT\bar{F}_{+-}=-2iT breaks the gauge group SU(3){\rm SU}(3) to U(1)×U(1){\rm U}(1)\times{\rm U}(1). The TT-charge of the components of ϕ+\phi_{+} is then

q=[012101210].q\ =\ \left[\begin{array}[]{ccc}0&1&2\\ -1&0&1\\ -2&-1&0\end{array}\right]. (4.53)

Therefore, there are two symmetric Higgs fields which are given by

ϕ+=[0φ1000φ2000],ϕ=[000φ1000φ20].\phi_{+}\ =\ \left[\begin{array}[]{ccc}0&\varphi_{1}&0\\ 0&0&\varphi_{2}\\ 0&0&0\end{array}\right],\hskip 28.45274pt\phi_{-}\ =\ \left[\begin{array}[]{ccc}0&0&0\\ \varphi_{1}^{\dagger}&0&0\\ 0&\varphi_{2}^{\dagger}&0\end{array}\right]. (4.54)

This is our first example where two independent symmetric Higgs fields appear. In the previous section, we also have two symmetric Higgs fields, but they form a doublet of SU(2){\rm SU}(2) gauge group.

The scalar potential for this case becomes

V(ϕ)=18(|φ1|22)2+18(|φ2|2|φ1|2)2+18(|φ2|22)2.V(\phi)\ =\frac{1}{8}\left(|\varphi_{1}|^{2}-2\right)^{2}+\frac{1}{8}\left(|\varphi_{2}|^{2}-|\varphi_{1}|^{2}\right)^{2}+\frac{1}{8}\left(|\varphi_{2}|^{2}-2\right)^{2}. (4.55)

The Higgs vacuum corresponds to φ1=φ2=2\varphi_{1}=\varphi_{2}=\sqrt{2} up to U(1)×U(1){\rm U}(1)\times{\rm U}(1) transformation. This attains the global minimum of the scalar potential which implies that the original SU(3){\rm SU}(3) gauge symmetry should be recovered. Then at the minimum, the symmetric Higgs field ϕ+\phi_{+} has the vev given by

ϕ+=T+(1):=[020002000].\phi_{+}\ =T^{(1)}_{+}:=\left[\begin{array}[]{ccc}0&\sqrt{2}&0\\ 0&0&\sqrt{2}\\ 0&0&0\end{array}\right]. (4.56)

This is the spin-1 representation T+(1)T^{(1)}_{+} of t+t_{+} embedded into su(3)su(3). Then the counting in this case is also reduced to a group-theoretic calculation.

The massless condition (4.9), in which T+T_{+} is replaced with T+(1)T^{(1)}_{+}, can be written as

D¯+[aμ,11aμ,12aμ,13aμ,21aμ,22aμ,23aμ,31aμ,32aμ,33]\displaystyle\bar{D}_{+}\left[\begin{array}[]{ccc}a_{\mu,11}&a_{\mu,12}&a_{\mu,13}\\ a_{\mu,21}&a_{\mu,22}&a_{\mu,23}\\ a_{\mu,31}&a_{\mu,32}&a_{\mu,33}\end{array}\right] (4.60)
=\displaystyle= i[2aμ,212(aμ,22aμ,11)2(aμ,23aμ,12)2aμ,312(aμ,32aμ,21)2(aμ,33aμ,22)02aμ,312aμ,32].\displaystyle-i\left[\begin{array}[]{ccc}\sqrt{2}a_{\mu,21}&\sqrt{2}(a_{\mu,22}-a_{\mu,11})&\sqrt{2}(a_{\mu,23}-a_{\mu,12})\\ \sqrt{2}a_{\mu,31}&\sqrt{2}(a_{\mu,32}-a_{\mu,21})&\sqrt{2}(a_{\mu,33}-a_{\mu,22})\\ 0&-\sqrt{2}a_{\mu,31}&-\sqrt{2}a_{\mu,32}\end{array}\right]. (4.64)

This can be rearranged into two sets of equations. One is

D¯+[2aμ,13aμ,23aμ,1213(aμ,112aμ,22+aμ,33)aμ,21aμ,322aμ,31]\displaystyle\bar{D}_{+}\left[\begin{array}[]{c}\sqrt{2}a_{\mu,13}\\ a_{\mu,23}-a_{\mu,12}\\ \frac{1}{\sqrt{3}}(a_{\mu,11}-2a_{\mu,22}+a_{\mu,33})\\ a_{\mu,21}-a_{\mu,32}\\ \sqrt{2}a_{\mu,31}\end{array}\right] =\displaystyle= iT+(2)[2aμ,13aμ,23aμ,1213(aμ,112aμ,22+aμ,33)aμ,21aμ,322aμ,31],\displaystyle-iT^{(2)}_{+}\left[\begin{array}[]{c}\sqrt{2}a_{\mu,13}\\ a_{\mu,23}-a_{\mu,12}\\ \frac{1}{\sqrt{3}}(a_{\mu,11}-2a_{\mu,22}+a_{\mu,33})\\ a_{\mu,21}-a_{\mu,32}\\ \sqrt{2}a_{\mu,31}\end{array}\right], (4.75)

where

T+(2):=[0200000600000600000200000]T^{(2)}_{+}\ :=\left[\begin{array}[]{ccccc}0&2&0&0&0\\ 0&0&\sqrt{6}&0&0\\ 0&0&0&\sqrt{6}&0\\ 0&0&0&0&2\\ 0&0&0&0&0\end{array}\right] (4.77)

is the spin-2 representation of t+t_{+}. The other is

D¯+[aμ,23+aμ,12aμ,33aμ,11aμ,21aμ,32]\displaystyle\bar{D}_{+}\left[\begin{array}[]{c}a_{\mu,23}+a_{\mu,12}\\ a_{\mu,33}-a_{\mu,11}\\ -a_{\mu,21}-a_{\mu,32}\end{array}\right] =\displaystyle= iT+(1)[aμ,23+aμ,12aμ,33aμ,11aμ,21aμ,32],\displaystyle-iT^{(1)}_{+}\left[\begin{array}[]{c}a_{\mu,23}+a_{\mu,12}\\ a_{\mu,33}-a_{\mu,11}\\ -a_{\mu,21}-a_{\mu,32}\end{array}\right], (4.84)

where T+(1)T^{(1)}_{+} is the spin-1 representation of t+t_{+} given in Eq.(4.56). These two sets of equations correspond to the irreducible decomposition of 𝟖\bf 8 of su(3)su(3) into 𝟓𝟑\bf 5\oplus 3 of an su(2)su(2) subgroup.

Recall that the TT-charges for aμa_{\mu} are given also as (4.53). Each linear combination in the above equations consists of components with the same TT-charge, as it should be. We find that the spin-2 representation in the mode expansion gives the solution for the condition (LABEL:condition_massless_vector_spin_2), and the spin-1 representation gives the solution for the condition (4.84). They give eight massless vector fields at the symmetric Higgs vacuum, and as expected, the symmetry pattern is given by

GYM=SU(3)backgroundfluxU(1)×U(1)HiggsvevSU(3).G_{\rm YM}={\rm SU}(3)\xrightarrow{{\rm background}\ {\rm flux}}{\rm U}(1)\times{\rm U}(1)\xrightarrow{{\rm Higgs}\ {\rm vev}}{\rm SU}(3)\,. (4.85)

5 Higgs condensation on 2=SU(3)/(SU(2)×U(1))\mathbb{CP}^{2}={\rm SU}(3)/({\rm SU}(2)\times{\rm U}(1))

In this section, we consider Yang-Mills theories on 4×2\mathbb{R}^{4}\times\mathbb{CP}^{2} with the gauge group SU(4). Since 2\mathbb{CP}^{2} can be represented as a coset space G/H=SU(3)/(SU(2)×U(1))G/H={\rm SU}(3)/({\rm SU}(2)\times{\rm U}(1)), we can apply to this case techniques similar to those in the previous section.

The choice of the H=SU(2)×U(1)H={\rm SU}(2)\times{\rm U}(1) subgroup of G=SU(3)G={\rm SU}(3) is specified by their generators given as

ta:=12[σa00000],t4:=13[100010002].t_{a}\ :=\ \frac{1}{2}\left[\begin{array}[]{cc}\hskip 8.53581pt\sigma_{a}&\begin{array}[]{c}0\\ 0\end{array}\\ \begin{array}[]{cc}0&0\end{array}&0\end{array}\right],\hskip 28.45274ptt_{4}\ :=\ \frac{1}{3}\left[\begin{array}[]{ccc}1&0&0\\ 0&1&0\\ 0&0&-2\end{array}\right]. (5.1)

where a=1,2,3a=1,2,3. We choose the other generators of su(3)su(3) as

tz:=[001000000],tw:=[000001000],ts¯:=(ts),t_{z}\ :=\ \left[\begin{array}[]{ccc}0&0&1\\ 0&0&0\\ 0&0&0\\ \end{array}\right],\hskip 28.45274ptt_{w}\ :=\ \left[\begin{array}[]{ccc}0&0&0\\ 0&0&1\\ 0&0&0\end{array}\right],\hskip 28.45274ptt_{\bar{s}}\ :=\ (t_{s})^{\dagger}, (5.2)

where s=z,ws=z,w. According to this choice, the tangent space index mm takes z,w,z¯z,w,\bar{z} and w¯\bar{w}.

In the following, we always embed su(2)su(2) part, t1,t2,t3t_{1},t_{2},t_{3}, into 𝔤YM=su(4)\mathfrak{g}_{\rm YM}=su(4) as

Ta:=[0ta000000].T_{a}\ :=\ \left[\begin{array}[]{cccc}&&&0\\ &t_{a}&&0\\ &&&0\\ 0&0&0&0\end{array}\right]. (5.3)

In the following, we will study two cases of embedding T4T_{4} charge of the u(1)u(1) generator, t4t_{4}, into su(4)su(4).

The scalar potential (3.3) becomes

V(ϕ)\displaystyle V(\phi) =\displaystyle= 18Tr|F¯zz¯+i[ϕz,ϕz¯]|2+18Tr|F¯ww¯+i[ϕw,ϕw¯]|2\displaystyle\frac{1}{8}{\rm Tr}\left|\bar{F}_{z\bar{z}}+i[\phi_{z},\phi_{\bar{z}}]\right|^{2}+\frac{1}{8}{\rm Tr}\left|\bar{F}_{w\bar{w}}+i[\phi_{w},\phi_{\bar{w}}]\right|^{2} (5.4)
+14Tr|F¯zw+i[ϕz,ϕw]|2+14Tr|F¯zw¯+i[ϕz,ϕw¯]|2.\displaystyle+\frac{1}{4}{\rm Tr}\left|\bar{F}_{zw}+i[\phi_{z},\phi_{w}]\right|^{2}+\frac{1}{4}{\rm Tr}\left|\bar{F}_{z\bar{w}}+i[\phi_{z},\phi_{\bar{w}}]\right|^{2}.

Recall that the background flux is given as

F¯mn=faTamn,\bar{F}_{mn}\ =\ f^{a}{}_{mn}T_{a}, (5.5)

where famnf^{a}{}_{mn} are the structure constants of 𝔤=su(3)\mathfrak{g}=su(3), not of 𝔤YM=su(4)\mathfrak{g}_{\rm YM}=su(4). In this case we find F¯zw=0\bar{F}_{zw}=0 since the above choice of the generators for su(3)su(3) in (5.1) and (5.2) gives fa=zw0f^{a}{}_{zw}=0.

Recall that the background gauge field A¯α\bar{A}_{\alpha} is defined as

A¯α=eαaTa.\bar{A}_{\alpha}\ =\ e^{a}_{\alpha}T_{a}. (5.6)

In the previous section, eαae^{a}_{\alpha} gives a monopole configuration on S2S^{2}. Similarly, on 2\mathbb{CP}^{2}, eαae^{a}_{\alpha} gives an instanton background. This can be deduced from the fact that eαae^{a}_{\alpha} also gives the spin connection on 2\mathbb{CP}^{2} as (2.19), and that the second Chern number of 2\mathbb{CP}^{2} is non-zero [45].

In fact, this can be checked easily since the flux can be given expilcitly. We consider

f¯mn:=fatamn\bar{f}_{mn}\ :=\ f^{a}{}_{mn}t_{a} (5.7)

as a flux of the SU(2)×U(1){\rm SU}(2)\times{\rm U}(1) gauge field on 2\mathbb{CP}^{2}. We notice that

f¯zz¯=i(t3+32t4),f¯ww¯=i(t3+32t4),f¯zw¯=i(t1+it2),f¯zw= 0\bar{f}_{z\bar{z}}\ =\ -i\left(t_{3}+\frac{3}{2}t_{4}\right),\hskip 14.22636pt\bar{f}_{w\bar{w}}\ =\ -i\left(-t_{3}+\frac{3}{2}t_{4}\right),\hskip 14.22636pt\bar{f}_{z\bar{w}}\ =\ -i\left(t_{1}+it_{2}\right),\hskip 14.22636pt\bar{f}_{zw}\ =\ 0 (5.8)

satisfy

f¯zz¯+f¯ww¯=3it4,f¯zw= 0,f¯z¯w¯= 0.\bar{f}_{z\bar{z}}+\bar{f}_{w\bar{w}}\ =\ -3i\,t_{4},\hskip 28.45274pt\bar{f}_{zw}\ =\ 0,\hskip 28.45274pt\bar{f}_{\bar{z}\bar{w}}\ =\ 0. (5.9)

Note that f¯zz¯+f¯ww¯\bar{f}_{z\bar{z}}+\bar{f}_{w\bar{w}} is nonvanishing for the u(1)u(1) part, and there is no su(2)su(2) part. Since the instanton equation

Fmn=12ϵmnklFkl,F_{mn}\ =\ -\frac{1}{2}\epsilon_{mnkl}F_{kl}, (5.10)

can be rewritten in terms of the complex coordinates as

Fzz¯+Fww¯= 0,Fzw= 0,Fz¯w¯= 0,F_{z\bar{z}}+F_{w\bar{w}}\ =\ 0,\hskip 28.45274ptF_{zw}\ =\ 0,\hskip 28.45274ptF_{\bar{z}\bar{w}}\ =\ 0, (5.11)

we find that the su(2)su(2) part of the flux f¯mn\bar{f}_{mn} satisfies these instanton equation. Thus the su(2)su(2) field strength is given by an su(2)su(2) instanton configuration. But it does not always mean that the SU(4)\rm SU(4) configuration has a nonzero instanton number, as we will see later.

In the following, we consider two embeddings of SU(2)×U(1){\rm SU}(2)\times{\rm U}(1) into SU(4){\rm SU}(4), and investigate the corresponding Higgs condensations. We will see that the topological nature of the background SU(4){\rm SU}(4) flux plays an important role in the gauge symmetry pattern when the symmetric Higgs acquires vev.

5.1 Embedding of HH==SU(2)×(2)\timesU(1)(1) into GYMG_{\rm YM}==SU(4)(4): Case 1

Since SU(2)\rm SU(2) part of H=SU(2)×U(1)H={\rm SU(2)\times U(1)} is embedded into SU(4)\rm SU(4) as (5.3), we choose an embedding of U(1)H{\rm U(1)}\subset H part. Our first choice for T4u(1)T_{4}\in u(1) is

T4=13[1000010000200000].T_{4}\ =\ \frac{1}{3}\left[\begin{array}[]{cccc}1&0&0&0\\ 0&1&0&0\\ 0&0&-2&0\\ 0&0&0&0\end{array}\right]. (5.12)

The background flux F¯mn\bar{F}_{mn} breaks the SU(4){\rm SU}(4) gauge group to U(1)×U(1){\rm U}(1)\times{\rm U}(1).

Recall that the condition (3.5) for the symmetric Higgs fields is

[Ta,ϕm]=ifnϕnam[T_{a},\phi_{m}]\ =\ if^{n}{}_{am}\phi_{n} (5.13)

where TaT_{a} are Lie algebra generators of HH. In the previous section, this is a condition for a charge assigned to the components of ϕm\phi_{m}. For the 2\mathbb{CP}^{2} case, TaT_{a} form an su(2)×u(1)su(2)\times u(1) subalgebra of su(4)su(4), and the adjoint representation adYM{\rm ad}_{\rm YM}, i.e., 𝟏𝟓\bf 15 representation of su(4)su(4), can be decomposed into irreducible representations of su(2)×u(1)su(2)\times u(1) as

𝟏𝟓= 3𝟎𝟐𝟏𝟐𝟏𝟑𝟐𝟏𝟐𝟏𝟑𝟏𝟐𝟑𝟏𝟐𝟑(𝟏𝟎)2.\bf 15\ =\ 3_{0}\oplus 2_{1}\oplus 2_{\frac{1}{3}}\oplus 2_{-1}\oplus 2_{-\frac{1}{3}}\oplus 1_{\frac{2}{3}}\oplus 1_{-\frac{2}{3}}\oplus(1_{0})^{\rm 2}. (5.14)

On the other hand, the same commutation relations are realized by the original su(3)su(3) algebras,

[ta,tm]=ifntnam,[t_{a},t_{m}]\ =\ if^{n}{}_{am}t_{n}, (5.15)

and tmt_{m} forms a set of irreducible representations of su(2)×u(1)su(2)\times u(1):

Rt=𝟐𝟏𝟐𝟏.R_{t}=\bf 2_{1}\oplus 2_{-1}. (5.16)

Then the symmetric Higgs fields ϕm\phi_{m} satisfying (5.13) can be obtained by those representations of su(2)×u(1)su(2)\times u(1) isomorphic to RtR_{t} in the irreducible decomposition of the adjoint representation 𝟏𝟓\bf 15 of su(4)su(4). From the definitions of tmt_{m} in (5.2), we see that the symmetric Higgs fields are given by

ϕz=[00φ0000000000000],ϕw=[000000φ000000000],ϕs¯=(ϕs).\phi_{z}\ =\ \left[\begin{array}[]{cccc}0&0&\varphi&0\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&0\end{array}\right],\hskip 28.45274pt\phi_{w}\ =\ \left[\begin{array}[]{cccc}0&0&0&0\\ 0&0&\varphi&0\\ 0&0&0&0\\ 0&0&0&0\end{array}\right],\hskip 28.45274pt\phi_{\bar{s}}\ =\ (\phi_{s})^{\dagger}. (5.17)

Note that their non-zero components must be the same in order to satisfy (5.13).

The scalar potential V(ϕ)V(\phi) is then given by

V(ϕ)=34(1|φ|2)2.V(\phi)\ =\ \frac{3}{4}\left(1-|\varphi|^{2}\right)^{2}. (5.18)

Therefore, the symmetric Higgs field φ\varphi acquires vev at φ=1\varphi=1 up to a gauge transformation. This breaks the residual U(1)×U(1){\rm U}(1)\times{\rm U}(1) gauge symmetry to U(1){\rm U}(1), and within the analysis of the low lying modes in the coset space compactification, the symmetry breaking pattern would be

GYM=SU(4)backgroundfluxU(1)×U(1)HiggsvevU(1)?G_{\rm YM}={\rm SU}(4)\xrightarrow{{\rm background}\ {\rm flux}}{\rm U}(1)\times{\rm U}(1)\xrightarrow{{\rm Higgs}\ {\rm vev}}{\rm U}(1)\ ? (5.19)

We have found that the symmetric Higgs vacuum attains the global minimum of the scalar potential. This implies the stability of the vacuum, and the restoration of gauge symmetry when the Higgs acquires vev. In fact, this turns out to happen in more general situations [46, 47]. This will become apparent when we reconsider the above calculations as follows to elaborate the reason why |φ|=1|\varphi|=1 attains the global minimum of the potential in the present setup. We have considered the embedding (5.3) and (5.12), which can be generalized to the other generators as

Tm:=[0tm000000].T_{m}\ :=\ \left[\begin{array}[]{cccc}&&&0\\ &t_{m}&&0\\ &&&0\\ 0&0&0&0\end{array}\right]. (5.20)

Then, TaT_{a} and TmT_{m} form an su(3)su(3) subalgebra embedded into the 3×33\times 3 upper-left block of su(4)su(4). The condition for the symmetric Higgs fields is given by (5.13), and the condition for the vanishing scalar potential in (3.3) is written as

[ϕm,ϕn]=ifaTamn.[\phi_{m},\phi_{n}]\ =\ if^{a}{}_{mn}T_{a}. (5.21)

Comparing these two conditions, we find that ϕm=Tm\phi_{m}=T_{m} is a soution for both conditions since they become nothing but a part of the commutation relations of su(3)su(3). This is the reason why the symmetric Higgs vev φ=1\varphi=1 attains the global minimum of the scalar potential. Now, it is clear that this phenomenon always happens for a general coset space G/HG/H, if we choose an embedding of HH into GYMG_{\rm YM} which is induced by an embedding of GG into GYMG_{\rm YM}. In fact, we have already observed this phenomenon in section 4.1 for G/H=S2G/H=S^{2}.

At this point, one might be puzzled by the fact that the symmetric Higgs vacuum attains the global minimum of the scalar potential, especially when one remembers that the background gauge filed consists of an instanton configuration. On the one hand, the Higgs condensation is nothing but a continuous deformation of the gauge field configuration on 2\mathbb{CP}^{2}. On the other hand, the vanishing potential implies that the gauge field configuration is just a pure gauge. This looks contradicting to the topologically non-trivial nature of the instanton configuration. The resolution of this puzzle comes from the fact that the su(2)su(2) instanton is embedded into su(4)su(4) with a u(1)u(1) flux, and the instanton number is cancelled between the su(2)su(2) and u(1)u(1) parts. Indeed, we can calculate the instanton number of the background gauge field A¯α\bar{A}_{\alpha} for the su(4)su(4) gauge field explicitly, and find

12ϵmnklTrFmnFkl\displaystyle\frac{1}{2}\epsilon_{mnkl}{\rm Tr}\,F_{mn}F_{kl} =\displaystyle= Tr(Fzz¯Fww¯+FzwFz¯w¯Fzw¯Fz¯w)\displaystyle{\rm Tr}\left(-F_{z\bar{z}}F_{w\bar{w}}+F_{zw}F_{\bar{z}\bar{w}}-F_{z\bar{w}}F_{\bar{z}w}\right) (5.22)
=\displaystyle= Tr(T3T3+94T4T4T1T1T2T2)\displaystyle{\rm Tr}\left(-T_{3}T_{3}+\frac{9}{4}T_{4}T_{4}-T_{1}T_{1}-T_{2}T_{2}\right)
=\displaystyle= 0.\displaystyle 0.

The (T4)2(T_{4})^{2} part is a contribution from the u(1)u(1), and we conclude that the gauge field configuration before the Higgs condensation has zero instanton number. This is compatible with the fact that the gauge field configuration at the symmetric Higgs vacuum is trivial.

Finally, let us check whether there are 15 massless vector fields at the symmetric Higgs vacuum. The massless conditions in this case are

D¯zaμ+i[Tz,aμ]= 0,D¯waμ+i[Tw,aμ]= 0,\bar{D}_{z}a_{\mu}+i[T_{z},a_{\mu}]\ =\ 0,\hskip 28.45274pt\bar{D}_{w}a_{\mu}+i[T_{w},a_{\mu}]\ =\ 0, (5.23)

where TzT_{z} and TwT_{w} are defined in (5.20) with m=zm=z and ww, respectively. Since TaT_{a} and TmT_{m} form an su(3)su(3) subalgebra of su(4)su(4), as mentioned above, it is convenient to decompose these massless conditions according to the irreducible decomposition of 𝟏𝟓\bf 15 of su(4)su(4) to 𝟖𝟑𝟑¯𝟏\bf 8\oplus 3\oplus\bar{3}\oplus 1 of the su(3)su(3) subalgebra. By rearranging the components aμ,ija_{\mu,ij} into the corresponding vectors aμ(R)a^{(R)}_{\mu} with R=𝟖,𝟑,𝟑¯,𝟏R={\bf 8,3,\bar{3},1}, we obtain

D¯zaμ(R)=iTz(R)aμ(R),D¯waμ(R)=iTw(R)aμ(R),\bar{D}_{z}a^{(R)}_{\mu}\ =\ -iT^{(R)}_{z}\cdot a^{(R)}_{\mu},\hskip 28.45274pt\bar{D}_{w}a^{(R)}_{\mu}\ =\ -iT^{(R)}_{w}\cdot a^{(R)}_{\mu}, (5.24)

where Tm(R)T^{(R)}_{m} are generators in the representation RR.

As reviewed in Appendix F, each component of the vectors aμ(R)a^{(R)}_{\mu} on 2\mathbb{CP}^{2} can be expanded by the complete set of functions fIJR(y)f^{R}_{IJ}(y) where RR runs over all representations of SU(3){\rm SU}(3) and I,JI,J are the indices for the representation RR, that is, they run from 1 to dimR\dim R. Recall that the representation RR and one of the indices II are constrained by a condition of what kind of representation of HH we are investigating on the coset space G/HG/H. In the previous section, we used one index mm of the mode functions fmmj(y)f^{j}_{mm^{\prime}}(y) to indicate its TT-charges qq, and jj is constrained so that the representation contains the desired value of m=q.m=q. Similarly, in this case, we use one index II of fIJR(y)f^{R}_{IJ}(y) to indicate its su(2)×u(1)su(2)\times u(1) representation. Namely, if the irreducible decomposition of RR has a representation rr of su(2)×u(1)su(2)\times u(1), then fiJR(y)f^{R}_{iJ}(y) contributes to the expansion of a field in the representation rr of su(2)×u(1)su(2)\times u(1), where ii runs from 1 to dimr\dim r. Therefore, the expansion of a field χi(y)\chi_{i}(y) in the representation rr is given as

χi(y)=rRJ=1dimRcJRfiJR(y),\chi_{i}(y)\ =\ \sum_{r\subset R}\sum_{J=1}^{\dim R}c^{R}_{J}f^{R}_{iJ}(y), (5.25)

where the first summation is over the representations RR of su(3)su(3) whose irreducible decomposition with respect to su(2)×u(1)su(2)\times u(1) has rr. If the decomposition of RR contains several irreducibe representations each of which is isomorphic to rr, then the multiplicity is also taken into account in the sum.

The su(2)×u(1)su(2)\times u(1) representations for aμ(R)a^{(R)}_{\mu} can be found by further decomposition of RR with respect to su(2)×u(1)su(2)\times u(1) subalgebra of su(3)su(3). Explicitly,

𝟖\displaystyle\bf 8 =\displaystyle= 𝟑𝟎𝟐𝟏𝟐𝟏𝟏𝟎,\displaystyle\bf 3_{0}\oplus 2_{1}\oplus 2_{-1}\oplus 1_{0}, (5.26)
𝟑\displaystyle\bf 3 =\displaystyle= 𝟐𝟏𝟑𝟏𝟐𝟑,\displaystyle\bf 2_{\frac{1}{3}}\oplus 1_{-\frac{2}{3}}, (5.27)
𝟑¯\displaystyle\bf\bar{3} =\displaystyle= 𝟐𝟏𝟑𝟏𝟐𝟑,\displaystyle\bf 2_{-\frac{1}{3}}\oplus 1_{\frac{2}{3}}, (5.28)
𝟏\displaystyle\bf 1 =\displaystyle= 𝟏𝟎.\displaystyle\bf 1_{0}. (5.29)

The action of the covariant derivatives D¯z,D¯w\bar{D}_{z},\bar{D}_{w} on the mode functions fiJR(y)f^{R}_{iJ}(y) is again given by the multipication of Tz(R),Tw(R)T^{(R)}_{z},T^{(R)}_{w} from the left. Therefore, the massless condition is again reduced to the requirement that the adjoint action of Tz,TwT_{z},T_{w} due to the symmetric Higgs vev has the same effect on aμa_{\mu} as the action of Tz(R),Tw(R)T^{(R)}_{z},T^{(R)}_{w} on the mode functions.

In the present case, the solution to the massless conditions is almost obvious. For example, let i1,i2i_{1},i_{2} be indices for the representations 𝟐𝟏𝟑,𝟏𝟐𝟑\bf 2_{\frac{1}{3}},1_{-\frac{2}{3}}, respectively. Then

aμ(𝟑)(x,y)=[J=13cJ𝟑(x)fi1J𝟑(y)J=13cJ𝟑(x)fi2J𝟑(y)]a^{({\bf 3})}_{\mu}(x,y)\ =\ \left[\begin{array}[]{c}\sum_{J=1}^{3}c^{\bf 3}_{J}(x)f^{\bf 3}_{i_{1}J}(y)\\[5.69054pt] \sum_{J=1}^{3}c^{\bf 3}_{J}(x)f^{\bf 3}_{i_{2}J}(y)\end{array}\right] (5.30)

is the solution for R=𝟑R={\bf 3}. Note that the expansion coefficients in the first and the second rows are the same. The solutions for the other RR can be obtained similarly. They give us 15 massless vector fields, as expected. Thus, the gauge symmetry is restored and we have the symmetry breaking-restoration pattern

GYM=SU(4)backgroundfluxU(1)×U(1)HiggsvevSU(4).G_{\rm YM}={\rm SU}(4)\xrightarrow{{\rm background}\ {\rm flux}}{\rm U}(1)\times{\rm U}(1)\xrightarrow{{\rm Higgs}\ {\rm vev}}{\rm SU}(4).\ (5.31)

5.2 Embedding of HH==SU(2)×(2)\timesU(1)(1) into GYMG_{\rm YM}==SU(4)(4): Case 2

Let us consider a different embedding of HH into GYMG_{\rm YM}. The SU(2) part of H=SU(2)×U(1)H={\rm SU(2)\times U(1)} is embedded into GYM=SU(4)G_{\rm YM}={\rm SU(4)} as (5.3). Our second choice of the U(1) part, T4T_{4}, into SU(4)\rm SU(4) is

T4=12[1000010000100001].T_{4}\ =\ \frac{1}{2}\left[\begin{array}[]{cccc}1&0&0&0\\ 0&1&0&0\\ 0&0&-1&0\\ 0&0&0&-1\end{array}\right]. (5.32)

The corresponding background flux breaks the SU(4){\rm SU}(4) gauge group to SU(2)×U(1){\rm SU}(2)\times{\rm U}(1). The symmetric Higgs fields are then given by

ϕz=[00φ1φ2000000000000],ϕw=[000000φ1φ200000000].\phi_{z}\ =\ \left[\begin{array}[]{cccc}0&0&\varphi_{1}&\varphi_{2}\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&0\end{array}\right],\hskip 28.45274pt\phi_{w}\ =\ \left[\begin{array}[]{cccc}0&0&0&0\\ 0&0&\varphi_{1}&\varphi_{2}\\ 0&0&0&0\\ 0&0&0&0\end{array}\right]. (5.33)

Note that the irreducible decomposition of 𝟏𝟓\bf 15 is now

𝟏𝟓= 3𝟎(𝟐𝟏)2(𝟐𝟏)2(𝟏𝟎)4\bf 15\ =\ 3_{0}\oplus(2_{1})^{\rm 2}\oplus(2_{-1})^{\rm 2}\oplus(1_{0})^{\rm 4} (5.34)

due to the different choice of T4T_{4}. Thus there are two Rt=𝟐𝟏𝟐𝟏R_{t}=\bf 2_{1}\oplus 2_{-1} representations in 𝟏𝟓\bf 15, and we have two symmetric Higgs. The components φs\varphi_{s} (s=1,2)(s=1,2) form a doublet of the SU(2){\rm SU}(2) gauge group.

The calculation of the scalar potential for φs\varphi_{s} is rather complicated if we just insert the above expressions into (5.4). It is better to keep track of the residual gauge invariance. For this purpose, we rewrite a part of the commutation relations of su(4)su(4) relevant for calculating V(ϕ)V(\phi) so that the residual SU(2){\rm SU}(2) gauge symmetry becomes manifest.

Let us explicitly write some generators of su(4)su(4) other than TaT^{a}. First, we define

T~i:=[00000000000012σi].\tilde{T}_{i}\ :=\ \left[\begin{array}[]{cc}\begin{array}[]{cc}0&0\\ 0&0\end{array}&\begin{array}[]{cc}0&0\\ 0&0\end{array}\\ \begin{array}[]{cc}0&0\\ 0&0\end{array}&\frac{1}{2}\sigma_{i}\end{array}\right]. (5.35)

These three T~i\tilde{T}_{i} and T4T_{4} correspond to the generators of the residual gauge symmetry SU(2)×\timesU(1). The generators in the off-diagonal components are relabeled as

Tz1:=Tz,Tw1:=Tw,Tz2:=[0001000000000000],Tw2:=[0000000100000000],T_{z}^{1}\ :=\ T_{z},\hskip 14.22636ptT_{w}^{1}\ :=\ T_{w},\hskip 14.22636ptT_{z}^{2}\ :=\ \left[\begin{array}[]{cccc}0&0&0&1\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&0\end{array}\right],\hskip 14.22636ptT_{w}^{2}\ :=\ \left[\begin{array}[]{cccc}0&0&0&0\\ 0&0&0&1\\ 0&0&0&0\\ 0&0&0&0\end{array}\right], (5.36)

where TzT_{z} and TwT_{w} are defined in (5.20) with m=zm=z and ww, respectively, and Ts¯α:=(Tsα)T_{\bar{s}}^{\alpha}:=(T_{s}^{\alpha})^{\dagger} with s=z,ws=z,w. Then, the symmetric Higgs field ϕs\phi_{s} in (5.33) can be written in terms of these generators as

ϕs=α=1,2φαTsα,ϕs¯=α=1,2φαTs¯α.\phi_{s}\ =\sum_{\alpha=1,2}\varphi_{\alpha}T_{s}^{\alpha},\hskip 28.45274pt\phi_{\bar{s}}\ =\sum_{\alpha=1,2}\varphi_{\alpha}^{\dagger}T_{\bar{s}}^{\alpha}. (5.37)

The relevant commutation relations for calculating V(ϕ)V(\phi) are

[Tsα,Tt¯β]=(σi)t¯sδαβTi+δt¯sδαβT4δt¯s(σi)αβT~i,[T_{s}^{\alpha},T_{\bar{t}}^{\beta}]\ =\ (\sigma_{i})_{\bar{t}s}\delta^{\alpha\beta}T_{i}+\delta_{\bar{t}s}\delta^{\alpha\beta}T_{4}-\delta_{\bar{t}s}(\sigma_{i})^{\alpha\beta}\tilde{T}_{i}, (5.38)

where (σi)z¯z:=(σi)11(\sigma_{i})_{\bar{z}z}:=(\sigma_{i})_{11} etc. Then we obtain

[ϕs,ϕt¯]=|φ|2(σi)t¯sTi+|φ|2δt¯sT4δt¯s(φσiφ)T~i,[\phi_{s},\phi_{\bar{t}}]\ =\ |\varphi|^{2}(\sigma_{i})_{\bar{t}s}T_{i}+|\varphi|^{2}\delta_{\bar{t}s}T_{4}-\delta_{\bar{t}s}(\varphi\sigma_{i}\varphi^{\dagger})\tilde{T}_{i}, (5.39)

where |φ|2:=|φ1|2+|φ2|2|\varphi|^{2}:=|\varphi_{1}|^{2}+|\varphi_{2}|^{2}.

We also need to rewrite the commutation relation for tat_{a} in su(3)su(3) as

[ts,tt¯]=(σi)t¯sti+32δt¯st4.[t_{s},t_{\bar{t}}]\ =\ (\sigma_{i})_{\bar{t}s}t_{i}+\frac{3}{2}\delta_{\bar{t}s}t_{4}. (5.40)

Then the background flux can be written as

F¯st¯=i[(σi)t¯sTi+32δt¯sT4].\bar{F}_{s\bar{t}}\ =\ -i\left[(\sigma_{i})_{\bar{t}s}T_{i}+\frac{3}{2}\delta_{\bar{t}s}T_{4}\right]. (5.41)

Note that F¯st\bar{F}_{st} and F¯s¯t¯\bar{F}_{\bar{s}\bar{t}} vanish.

By using the above expressions, we find

F¯zz¯+i[ϕz,ϕz¯]\displaystyle\bar{F}_{z\bar{z}}+i[\phi_{z},\phi_{\bar{z}}] =\displaystyle= i(|φ|21)T3+i(|φ|232)T4i(φσiφ)T~i,\displaystyle i(|\varphi|^{2}-1)T_{3}+i\left(|\varphi|^{2}-\frac{3}{2}\right)T_{4}-i(\varphi\sigma_{i}\varphi^{\dagger})\tilde{T}_{i}, (5.42)
F¯ww¯+i[ϕw,ϕw¯]\displaystyle\bar{F}_{w\bar{w}}+i[\phi_{w},\phi_{\bar{w}}] =\displaystyle= i(|φ|21)T3+i(|φ|232)T4i(φσiφ)T~i,\displaystyle-i(|\varphi|^{2}-1)T_{3}+i\left(|\varphi|^{2}-\frac{3}{2}\right)T_{4}-i(\varphi\sigma_{i}\varphi^{\dagger})\tilde{T}_{i}, (5.43)
F¯zw¯+i[ϕz,ϕw¯]\displaystyle\bar{F}_{z\bar{w}}+i[\phi_{z},\phi_{\bar{w}}] =\displaystyle= i(|φ|21)(T1iT2).\displaystyle i(|\varphi|^{2}-1)(T_{1}-iT_{2}). (5.44)

Finally, the potential V(ϕ)V(\phi) turns out to be

V(ϕ)\displaystyle V(\phi) =\displaystyle= 14[12(1|φ|2)2+(32|φ|2)2+12(φσiφ)2]+14(1|φ|2)2\displaystyle\frac{1}{4}\left[\frac{1}{2}\left(1-|\varphi|^{2}\right)^{2}+\left(\frac{3}{2}-|\varphi|^{2}\right)^{2}+\frac{1}{2}(\varphi\sigma_{i}\varphi^{\dagger})^{2}\right]+\frac{1}{4}\left(1-|\varphi|^{2}\right)^{2} (5.45)
=\displaystyle= 34(1|φ|2)2+316.\displaystyle\frac{3}{4}\left(1-|\varphi|^{2}\right)^{2}+\frac{3}{16}.

The symmetric Higgs doublet therefore acquires the vev

φ1= 1,φ2= 0,\varphi_{1}\ =\ 1,\hskip 28.45274pt\varphi_{2}\ =\ 0, (5.46)

up to a global gauge transformation. The non-zero value of the scalar potential at the symmetric Higgs vacuum implies that there remains a flux at this vacuum, which suggests that the gauge symmetry at the symmetric Higgs vacuum must be smaller than SU(4){\rm SU}(4).

Now, we count the number of massless vector fields at the symmetric Higgs vacuum. The massless conditions are given as (5.23), exactly the same condition we discussed in the previous section. Therefore, we can employ the decomposition (5.24) again.

Let us consider

D¯zaμ(𝟑)=iTz(𝟑)aμ(𝟑),D¯waμ(𝟑)=iTw(𝟑)aμ(𝟑).\bar{D}_{z}a^{(\bf 3)}_{\mu}\ =\ -iT^{(\bf 3)}_{z}\cdot a^{(\bf 3)}_{\mu},\hskip 28.45274pt\bar{D}_{w}a^{(\bf 3)}_{\mu}\ =\ -iT^{(\bf 3)}_{w}\cdot a^{(\bf 3)}_{\mu}. (5.47)

Recall that aμ(𝟑)a^{(\bf 3)}_{\mu} is formed from the components aμ,14a_{\mu,14}, aμ,24a_{\mu,24} and aμ,34a_{\mu,34}. In terms of the su(2)×u(1)su(2)\times u(1) subgroup generated by TaT_{a}, this consists of the representations 𝟐𝟏𝟏𝟎\bf 2_{1}\oplus 1_{0}. As explained in Appendix F, they are expanded by the mode functions fIJR(y)f^{R}_{IJ}(y), where RR is a representation of su(3)su(3) whose irreducible decomposition contains 𝟐𝟏\bf 2_{1} or 𝟏𝟎\bf 1_{0}. An important point is that this irreducible decomposition must be considered with respect to the su(2)×u(1)su(2)\times u(1) subgroup (5.1), not to any su(4)su(4) embeddings. According to this, the representation 𝟑\bf 3 of the su(3)su(3) is decomposed as 𝟐𝟏𝟑𝟏𝟐𝟑\bf 2_{\frac{1}{3}}\oplus 1_{-\frac{2}{3}}. This implies that the mode functions fIJ𝟑(y)f^{\bf 3}_{IJ}(y) do not contribute to the expansion of aμ(𝟑)a^{(\bf 3)}_{\mu}. Instead, other mode functions, for example, fIJ𝟖(y)f^{\bf 8}_{IJ}(y) contribute to the expansion since 𝟖\bf 8 is decomposed as 𝟑𝟎𝟐𝟏𝟐𝟏𝟏𝟎\bf 3_{0}\oplus 2_{1}\oplus 2_{-1}\oplus 1_{0}. Then, the covariant derivatives D¯z,D¯w\bar{D}_{z},\bar{D}_{w} can be converted to Tz(𝟖),Tw(𝟖)T^{(\bf 8)}_{z},T^{(\bf 8)}_{w} but never to Tz(𝟑),Tw(𝟑)T^{(\bf 3)}_{z},T^{(\bf 3)}_{w}. By this reason, we conclude that the massless conditions (5.47) do not have solutions. Similarly, the conjugate components aμ(𝟑¯)a^{(\bf\bar{3})}_{\mu} give us only massive vector fields.

The conditions for the other two representations 𝟖\bf 8 and 𝟏\bf 1 turn out to give us 8+18+1 massless vector fields. A natural guess is that the emergent gauge symmetry at the symmetric Higgs vacuum would be SU(3)×U(1)\rm SU(3)\times U(1).

GYM=SU(4)backgroundfluxSU(2)×U(1)HiggsvevSU(3)×U(1).G_{\rm YM}={\rm SU}(4)\xrightarrow{{\rm background}\ {\rm flux}}{\rm SU}(2)\times{\rm U}(1)\xrightarrow{{\rm Higgs}\ {\rm vev}}{\rm SU}(3)\times{\rm U}(1).\ (5.48)

This should be confirmed by a further analysis.

At the symmetric Higgs vacuum obtained above, the scalar potential is non-vanishing. This is a similar situation as the one discussed in section 4.2. Interestingly, we can show that the symmetric Higgs vacuum in this section is stable, and the topological nature of the background gauge field plays an important role. In the following, we see that the background gauge field A¯α\bar{A}_{\alpha} before the Higgs condensation has a non-zero instanton number, which is unchanged by any continuous deformation of the gauge field. The non-zero instanton number is obtained in this setup since the embedding T4T_{4} of t4t_{4} is different from the one we chose in the previous section. Let us calculate the instanton number for the gauge field configuration, including the symmetric Higgs vev, given by

Fmn=F¯mn+i[ϕm,ϕn],ϕm=φαTmα.F_{mn}\ =\ \bar{F}_{mn}+i[\phi_{m},\phi_{n}],\hskip 28.45274pt\phi_{m}\ =\ \varphi_{\alpha}T^{\alpha}_{m}. (5.49)

Their explicit forms are given in (5.42)(5.43)(5.44). We find

12ϵmnklTrFmnFkl\displaystyle\frac{1}{2}\epsilon_{mnkl}{\rm Tr}\,F_{mn}F_{kl} =\displaystyle= 12(|φ|21)2(|φ|232)212(φσiφ)2+(|φ|21)2\displaystyle\frac{1}{2}\left(|\varphi|^{2}-1\right)^{2}-\left(|\varphi|^{2}-\frac{3}{2}\right)^{2}-\frac{1}{2}\left(\varphi\sigma_{i}\varphi^{\dagger}\right)^{2}+\left(|\varphi|^{2}-1\right)^{2} (5.50)
=\displaystyle= 34.\displaystyle-\frac{3}{4}.

The scalar potential is bounded by this instanton number density as

V(ϕ)\displaystyle V(\phi) =\displaystyle= 18Tr(Fmn+12ϵmnklFkl)218ϵmnklTrFmnFkl\displaystyle\frac{1}{8}{\rm Tr}\,\left(F_{mn}+\frac{1}{2}\epsilon_{mnkl}F_{kl}\right)^{2}-\frac{1}{8}\epsilon_{mnkl}{\rm Tr}\,F_{mn}F_{kl} (5.51)
\displaystyle\geq 316.\displaystyle\frac{3}{16}.

This shows that the symmetric Higgs vacuum attains the global minimum of the scalar potential in a given topological sector of the gauge field. In other words, we can say that the non-trivial topological nature of the original background gauge field A¯α\bar{A}_{\alpha} stabilizes the non-trivial symmetric Higgs vacuum.

This example tells us the geometric picture of the Higgs condensation realized in the Kaluza-Klein reduction of Yang-Mills theory. Since the Higgs fields come from some components of the gauge field, the Higgs condensation is nothing but a particular continuous deformation of the gauge field. The condensation occurs in order to minimize the “Euclidean action”

SE:=14𝑑vTrFmnFmn.S_{E}\ :=\ \frac{1}{4}\int dv\,{\rm Tr}\,F_{mn}F_{mn}. (5.52)

The configuration space of the gauge field is divided into various components according to certain topological invariants. The vacuum corresponds to the global minimum of SES_{E} in a given component. The global minimum may or may not be attained by the condensation of symmetric Higgs fields, according to the situation.

It is interesting to notice that the gauge field configuration FmnF_{mn} including the symmetric Higgs vev becomes exactly an instanton configuration at the symmetric Higgs vacuum. This means that we find an explicit construction of the instanton solution on 2\mathbb{CP}^{2}. It is interesting to clarify how general this construction is.

6 Conclusions

Higher-dimensional gauge theories with non-trivial fluxes of the background gauge fields in the compact space have been investigated as phenomenological models of gauge symmetry breaking. The background gauge fluxes explicitly break some of the original gauge symmetries, and often provide tachyonic scalar fields whose vacuum expectation value further breaks the remaining gauge symmetries in the four dimensional effective field theory. Thus, this kind of models are often utilized for dynamical generation of Higgs potential. In their investigations, we usually study four-dimensional effective field theories by keeping only the light fields and neglecting other massive modes in the Kaluza-Klein tower, since we are interested in physics in the lower energy scale than that of the compact space.

In this paper, we revisit such higher dimensional models including all the massive Kaluza-Klein modes to investigate their roles in the gauge symmetry breaking. The inclusion of higher modes will be important since the scale of the vacuum expectation value of the Higgs field is usually given by the same scale as the mass of the Kaluza-Klein modes. Indeed both of them are given by the scale of the compact space. What we have shown in the present paper is that, when the Higgs acquires vacuum expectation value at the minimum of the potential, some of the originally massive vector fields in the Kaluza-Klein tower become massless and the corresponding gauge symmetries are restored. If we restricted ourselves to consider only the light modes, we would conclude that the gauge symmetry, which remains to be unbroken by the gauge flux, would be further broken by the Higgs field. But, if we consider all the Kaluza-Klein modes, the symmetries, on the contrary, are restored even to the full set of the original gauge symmetries. When the massive gauge field becomes massless, it will provide an additional massless scalar field to the massless gauge field. Possible candidates for such massless scalar fields are, for example, instanton moduli in the model discussed in section 5.2. It is interesting to see whether the scalar field acquires mass due to the radiative corrections or remains massless against perturbations.

We have studied two classes of the compact space, S2S^{2} and 2\mathbb{CP}^{2}. Both of them are coset spaces G/HG/H, and due to the beautiful group theoretical structures of the coset space, we have succeeded to analyze the mass spectrum even when the Higgs acquires vacuum expectation value. In some cases, all the gauge symmetries are recovered, and in other cases, only a part of them are recovered. The vacuum is shown to be stable even in the latter cases. The stability is related to the topological structures of the gauge field configurations. In cases when the background gauge field configurations are topologically trivial, the original gauge symmetries are completely restored at the global minimum of the potential. On the other hand, in cases when the background gauge configurations are topologically non-trivial, or when they have some conserved topological numbers, the original gauge symmetries can be only partially restored. The background gauge field configurations including the Higgs fields in the true vacua are described by new topologically non-trivial configurations with the same topological numbers.

We have developed a group-theoretic technique for analyzing the number of massless vector fields at the symmetric Higgs vacuum. It is reasonalble to expect that the technique can be extended to obtain the mass formulas applicable also to massive vector fields and scalar fields. A key result which enables the technique to work is that the symmetric Higgs vacuum expectation values coincide with some generators of the gauge group GYMG_{\rm YM}. It is interesting to clarify whether this happens in more general models. A detailed understanding of the structure of the scalar potential will help us to gain insights on this issue.

An interesting observation we made is that the Higgs vacuum in a model invetigated in section 5.2 corresponds to an instanton solution on 2\mathbb{CP}^{2}. The solution can be given quite explicitly from the Maurer-Cartan 1-form on GG. It is curious to see whether this can be a general method of constructing instanton solutions on coset spaces. Our calculation suggests that the different embeddings of HH into GYMG_{\rm YM} would give us instanton solutions with different instanton numbers.

The effects of gravity are neglected in the investigations of the present paper. One of the original physics target of the coset compactifications is the stabilization of the compact space with gravity, called, “spontaneous compactification”, in which the background flux in the coset space is a classical solution of the gauge field equations as well as the Einstein equation. It will be interesting to investigate the stability and the pattern of gauge symmetry breaking including the fluctuations of gravity.

The Einstein-Yang-Mills theory naturally appears as the bosonic part of the low energy effective theory of the heterotic string theory. A Higgs field which appears in this context corresponds to a closed string tachyon, whose condensation is an interesting research subject in theoretically as well as phenomenologically. The condensation of closed string tachyon was discussed, for example, in [48, 49]. Since the Einstein-Yang-Mills theory can be regarded as a truncation of string field theory, it is a natural arena for discussing the Higgs (closed string tachyon) condensation. It is interesting if there would exist an endpoint of the condensation which is stabilized due to a topological reason.

Acknowledgements

The work is supported in part by Grants-in-Aid for Scientific Research No.16K05329, No.18H03708 and No.19K03851 from the Japan Society for the Promotion of Science.

Appendix A Coset spaces

In this appendix, we summarize some mathematical basics of coset spaces. For further details, see, e.g., [25] or [50].

Let GG be a group, and HH be a subgroup of GG. For an element gg of GG, we define a subset gHgH of GG as

gH:={gh|hH}.gH\ :=\ \{\,gh\,|\,h\in H\,\}. (A.1)

The set of such subsets is denoted by G/HG/H.

Suppose that gg belongs to both g1Hg_{1}H and g2Hg_{2}H in G/HG/H. Then, gg can be written as g=g1h1g=g_{1}h_{1} and g=g2h2g=g_{2}h_{2}, which imply g1=g2h2h11g_{1}=g_{2}h_{2}h_{1}^{-1}. This then implies g1H=g2Hg_{1}H=g_{2}H. In other words, g1Hg_{1}H and g2Hg_{2}H in G/HG/H are either equal or disjoint as subsets of GG.

We choose one representation giHg_{i}H for each element of G/HG/H. Then, the set {gi}\{g_{i}\} of elements of GG is in one-to-one correspondence to G/HG/H. The elements gig_{i} are called a set of representatives for G/HG/H. In terms of the representatives, GG can be written as

G=igiH,G\ =\ \bigcup_{i}g_{i}H, (A.2)

where giHg_{i}H and gjHg_{j}H are disjoint to each other if gigjg_{i}\neq g_{j}.

We can perform the same construction when GG is a Lie group and HH is its closed subgroup. Note that HH is always a Lie subgroup of GG. In this case, G/HG/H is known to be a smooth manifold. This is called a coset space. In mathematics literature, this is also called a homogeneous space. If the dimensions of GG and HH are dGd_{G} and dHd_{H}, respectively, then the dimension of G/HG/H is d:=dGdHd:=d_{G}-d_{H}. A set of representatives for G/HG/H is regarded as a GG-valued function on G/HG/H. For a given local coordinate patch UU of G/HG/H, the representatives corresponding to points in UU can be chosen such that they are given by a smooth function g(y)g(y) on UU where yαy^{\alpha} are coordinates on UU.

In this paper, we focus our attention on a particular class of coset spaces, known as symmetric coset spaces. They are characterized by the structures of the Lie algebras 𝔤\mathfrak{g} and 𝔥\mathfrak{h} of GG and HH, respectively. Let tat_{a} be a set of generators of 𝔥\mathfrak{h} where aa runs from 1 to dHd_{H}. Their commutation relations are

[ta,tb]=ifctcab,[t_{a},t_{b}]\ =\ if^{c}{}_{ab}t_{c}, (A.3)

since 𝔥\mathfrak{h} is a subalgebra of 𝔤\mathfrak{g}. Let tmt_{m} be a set of generators of 𝔤\mathfrak{g} other than tat_{a} where mm runs from 1 to dd. We require that the other commutation relations of 𝔤\mathfrak{g} are of the form

[ta,tm]=ifntnam,[tm,tn]=ifatamn.[t_{a},t_{m}]\ =\ if^{n}{}_{am}t_{n},\hskip 28.45274pt[t_{m},t_{n}]\ =\ if^{a}{}_{mn}t_{a}. (A.4)

A coset space G/HG/H in which the Lie algebras 𝔤,𝔥\mathfrak{g,h} satisfy this condition is said to be symmetric. This condition can be understood as follows. For a symmetric coset space, we can assign a “parity” for generators such that tat_{a} has +1+1 and tmt_{m} has 1-1. In sections 4 and 5, we consider particular symmetric coset spaces, namely S2S^{2} and 2\mathbb{CP}^{2}.

First, let us describe S2S^{2} as a symmetric coset space. For this purpose, we regard S2S^{2} as 1\mathbb{CP}^{1}. A point in 1\mathbb{CP}^{1} is represented by a pair of complex numbers (c1,c2)2(c_{1},c_{2})\in\mathbb{C}^{2} with (c1,c2)(0,0)(c_{1},c_{2})\neq(0,0). Two such pairs (c1,c2)(c_{1},c_{2}) and (c1,c2)(c_{1}^{\prime},c_{2}^{\prime}) correspond to the same point of 1\mathbb{CP}^{1} if and only if there exists a non-zero λ\lambda\in\mathbb{C} such that

(c1,c2)=(λc1,λc2)(c_{1}^{\prime},c_{2}^{\prime})\ =\ (\lambda c_{1},\lambda c_{2}) (A.5)

is satisfied. Using this ambiguity, we can always choose the pair (c1,c2)(c_{1},c_{2}) such that they satisfy |c1|2+|c2|2=1|c_{1}|^{2}+|c_{2}|^{2}=1. Any pair (c1,c2)(c_{1},c_{2}) satisfying |c1|2+|c2|2=1|c_{1}|^{2}+|c_{2}|^{2}=1 can be written as

[c1c2]=U[10],USU(2).\left[\begin{array}[]{c}c_{1}\\ c_{2}\end{array}\right]\ =\ U\left[\begin{array}[]{c}1\\ 0\end{array}\right],\hskip 28.45274ptU\in{\rm SU}(2). (A.6)

A pair (c1,c2)(c_{1},c_{2}) represents the same point in 1\mathbb{CP}^{1} as (1,0)(1,0) if and only if UU is of the form

U=[eiφ00eiφ].U\ =\ \left[\begin{array}[]{cc}e^{i\varphi}&0\\ 0&e^{-i\varphi}\end{array}\right]. (A.7)

Such elements form a U(1){\rm U}(1) subgroup of SU(2){\rm SU}(2). We have found that S2=1S^{2}=\mathbb{CP}^{1} can be written as the coset space SU(2)/U(1){\rm SU}(2)/{\rm U}(1). In this case, tat_{a} corresponds to 12σ3\frac{1}{2}\sigma_{3} and tmt_{m} correspond to σ±\sigma_{\pm}. Their commutation relations show that SU(2)/U(1){\rm SU}(2)/{\rm U}(1) is symmetric.

Next, consider 2\mathbb{CP}^{2}. This is a straightforward generalization of the 1\mathbb{CP}^{1} case. Any point of 2\mathbb{CP}^{2} corresponds to a triple (c1,c2,c3)3(c_{1},c_{2},c_{3})\in\mathbb{C}^{3} satisfying |c1|2+|c2|2+|c3|2=1|c_{1}|^{2}+|c_{2}|^{2}+|c_{3}|^{2}=1, which can be written as

[c1c2c3]=U[100],USU(3).\left[\begin{array}[]{c}c_{1}\\ c_{2}\\ c_{3}\end{array}\right]\ =\ U\left[\begin{array}[]{c}1\\ 0\\ 0\end{array}\right],\hskip 28.45274ptU\in{\rm SU}(3). (A.8)

The multiplication by UU does not change the point in 2\mathbb{CP}^{2} if and only if UU is of the form

[e2iφ0 000eiφU],USU(2).\left[\begin{array}[]{cc}e^{2i\varphi}&0\ \ 0\\ \begin{array}[]{c}0\\ 0\end{array}&e^{-i\varphi}U\\ \end{array}\right],\hskip 28.45274ptU\in\rm SU(2). (A.9)

The matrices of the form consist of an SU(2)×U(1)\rm SU(2)\times U(1) subgroup of SU(3)\rm SU(3). We have found that 2\mathbb{CP}^{2} can be written as a coset space SU(3)/(SU(2)×U(1))\rm SU(3)/(SU(2)\times U(1)). This is apparently a symmetric coset space.

Appendix B GG as a principal HH-bundle

In this paper, a coset space G/HG/H is used as a compactification manifold \cal M of a higher-dimensional Yang-Mills theory. Since we would like to define a gauge theory on G/HG/H with the gauge group GYMG_{\rm YM}, we need a principal GYMG_{\rm YM}-bundle on G/HG/H on which we can define the gauge field as a connection. Note that, in this paper, we discuss only classical aspects of Yang-Mills theory on a fixed principal GYMG_{\rm YM}-bundle, and the summation over different bundles will not be considered.

It is known [50] that there exists a natural principal HH-bundle on G/HG/H, as will be reviewed shortly. In fact, it is GG itself. This principal HH-bundle is specified by a set of transition functions gij(y){\rm g}_{ij}(y), whose values are in HH, defined on the overlap UiUjU_{i}\cap U_{j} of two local coordinate patches Ui,UjG/HU_{i},U_{j}\subset G/H. If we choose an embedding of HH into the gauge group GYMG_{\rm YM}, then gij(y){\rm g}_{ij}(y) can be regarded as a set of transition functions whose values are in GYMG_{\rm YM}. Using these transition functions, we can construct a principal GYMG_{\rm YM}-bundle on G/HG/H. In this manner, we can define Yang-Mills theory on G/HG/H whose gauge group is GYMG_{\rm YM}.

Recall the decomposition (A.2) of GG. This suggests that GG is a fiber bundle whose fibers are of the form gHgH. The subset gHgH is in one-to-one correspondence to HH for any gg. In fact, it can be shown that GG can be regarded as a fiber bundle whose fibers are HH, that is, a principal HH-bundle on G/HG/H. For the coset spaces, the representatives gig_{i} are replaced with g(y)g(y) chosen for each local coordinate patch. A choice of g(y)g(y) on UU amounts to choosing a local section g(y)g(y) on UU of the principal HH-bundle GG. The situation is depicted in Figure 1.

Refer to caption
Figure 1: A Lie group GG can be regarded as a principal HH-bundle on G/HG/H. The vertical direction corresponds to the fiber HH. A local section g(y)g(y) on UU is also depicted.

Let g(y)g^{\prime}(y) be another local section on UU. This is related to the original one by

g(y)=g(y)h(y),h(y)H.g^{\prime}(y)\ =\ g(y)h(y),\hskip 28.45274pth(y)\in H. (B.1)

Note that this transformation does not transform a point in UU to a different one since

g(y)H=g(y)h(y)H=g(y)H.g^{\prime}(y)H\ =\ g(y)h(y)H\ =\ g(y)H. (B.2)

Indeed, this amounts to a local gauge transformation with respect to h(y)HGYMh(y)\in H\subset G_{\rm YM}, as we will show in Appendix C. In fact, this also induces a local Lorentz transformation simultaneously. To avoid possible confusions with the ordinary gauge transformation, we call the local transformation induced by the right-multiplication of h(y)h(y) an HH-transformation.

We can multiply any yy-independent element g0g_{0} to the representatives g(y)g(y) from the left. The result g0g(y)g_{0}\cdot g(y) is also an element of GG, so this must be in a certain subset g(y)Hg(y^{\prime})H for some point yy^{\prime}. This can be written as

g0g(y)=g(y)h(y,g0).g_{0}\cdot g(y)\ =\ g(y^{\prime})h(y,g_{0}). (B.3)

In general, the pont yy^{\prime} is different from the original yy. Therefore, the left-multiplication of g0g_{0} induces a coordinate transformation of G/HG/H. In addition, this simultaneously induces an HH-transformation with respect to h(y,g0)h(y,g_{0}). We will see in the next section that this is actually an isometry of G/HG/H for a natural choice of the metric on G/HG/H. In fact, the isometry group of G/HG/H with respect to the metric is known to be isomorphic to GG, that is, any isometry of G/HG/H is induced by the left-multiplication as above.

Appendix C Maurer-Cartan 1-form

There exists a 1-form on GG which can be defined without any additional information. We review in the following that this 1-form defines a natural metric and background gauge field on G/HG/H. See [25] and [50] for details.

Let gij(ξ)g_{ij}(\xi) be a matrix-valued function on a local coordinate patch of GG whose value at a point ξG\xi\in G is an N×NN\times N matrix representing ξ\xi. This can be regarded as a set of coordinate functions on GG. Therefore, like dyαdy^{\alpha} on UG/HU\subset G/H, we have a set of functions dgij(ξ)dg_{ij}(\xi) on GG whose values are 1-forms on GG. Using them, we construct

(g1)ij(ξ)dgjk(ξ).(g^{-1})_{ij}(\xi)\,dg_{jk}(\xi). (C.1)

This is called the Maurer-Cartan 1-form on GG. In the following, we simply denote this by g1dgg^{-1}dg.

A local section g(y)g(y) can be regarded as an embedding of the local coordinate patch UU into GG. The pull-back of g1dgg^{-1}dg with respect to this embedding gives us a 1-form g(y)1dg(y)g(y)^{-1}dg(y) on UU. The latter can be expanded as

g(y)1dg(y)=iemtm+ieata,g(y)^{-1}dg(y)\ =\ ie^{m}t_{m}+ie^{a}t_{a}, (C.2)

where

em=eαmdyα,ea=eαadyα,e^{m}\ =\ e^{m}_{\alpha}dy^{\alpha},\hskip 28.45274pte^{a}\ =\ e^{a}_{\alpha}dy^{\alpha}, (C.3)

since g(y)1αg(y)g(y)^{-1}\partial_{\alpha}g(y) is in the Lie algebra 𝔤\mathfrak{g}. We will clarify the meanings of eαme^{m}_{\alpha} and eαae^{a}_{\alpha} in the following.

Recall that elements of the Lie algebra 𝔤\mathfrak{g} correspond to tangent vectors of GG at the identity element. For the coset space, tmt_{m} correspond to a set of basis of the tangent space of G/HG/H. Therefore, eαme^{m}_{\alpha} gives us a vielbein on G/HG/H. This is an invertible d×dd\times d matrix by construction.

To see the role played by eαae^{a}_{\alpha}, consider an HH-transformation g(y)g(y)h(y)g(y)\to g(y)h(y). This induces

g(y)1dg(y)h(y)1(g(y)1dg(y))h(y)+h(y)1dh(y).g(y)^{-1}dg(y)\ \to\ h(y)^{-1}\Bigl{(}g(y)^{-1}dg(y)\Bigr{)}h(y)+h(y)^{-1}dh(y). (C.4)

This implies

emh(y)1emh(y),eah(y)1eah(y)ih(y)1dh(y).e^{m}\ \to\ h(y)^{-1}e^{m}h(y),\hskip 28.45274pte^{a}\ \to\ h(y)^{-1}e^{a}h(y)-ih(y)^{-1}dh(y). (C.5)

Here, we have used the fact that

h(y)1tmh(y)𝔪h(y)^{-1}t_{m}h(y)\ \in\ \mathfrak{m} (C.6)

holds, where 𝔪\mathfrak{m} is a vector space spanned by tmt_{m}. This is due to the commutation relation

[ta,tm]=ifntnam[t_{a},t_{m}]\ =\ if^{n}{}_{am}t_{n} (C.7)

which we assumed in Appendix A. The transformation of eae^{a} shows that eαae^{a}_{\alpha} behaves as a gauge field on G/HG/H with the gauge group HH, defined on the principal HH-bundle GG.

Note that eme^{m} also transforms under the HH-transformation. Indeed, the equation (C.6) implies that this is a local Lorentz transformation acting on the tangent space of G/HG/H. Therefore, the right-multiplication of h(y)h(y) to g(y)g(y) induces both a gauge transformation of eαae^{a}_{\alpha} and a local Lorentz transformation of eαme^{m}_{\alpha} simultaneously.

Recall that the left-multiplication of g0Gg_{0}\in G to a local section g(y)g(y) induces a coordinate transformation yyy\to y^{\prime} of G/HG/H. The 1-form g(y)1dg(y)g(y)^{-1}dg(y) on UU is trivially invariant under g(y)g0g(y)g(y)\to g_{0}\cdot g(y) for a yy-independent g0g_{0}. This implies that the vielbein eαme^{m}_{\alpha} (and the gauge field eαae^{a}_{\alpha}) are invariant under the coordinate transformation. In other words, the coordinate transformation is an isometry with respect to the metric hαβh_{\alpha\beta} constructed from eαme^{m}_{\alpha}. It is known that any isometry of G/HG/H is induced in this manner. Therefore, the isometry group of G/HG/H is isomorphic to GG.

We calculate the field strength 2-form

12fata:=deata+ieataebtb=deata12tcfceaabeb.\frac{1}{2}f^{a}t_{a}\ :=\ de^{a}t_{a}+ie^{a}t_{a}\wedge e^{b}t_{b}\ =\ de^{a}t_{a}-\frac{1}{2}t_{c}f^{c}{}_{ab}e^{a}\wedge e^{b}. (C.8)

For this purpose, we do not need to know the explicit form of eαae^{a}_{\alpha}. Instead, we start with the following identity

d(g(y)1dg(y))=g(y)1dg(y)g(y)1dg(y).d\Bigl{(}g(y)^{-1}dg(y)\Bigr{)}\ =\ -g(y)^{-1}dg(y)\wedge g(y)^{-1}dg(y). (C.9)

This can be decomposed into the following two equations

deata\displaystyle de^{a}t_{a} =\displaystyle= 12tcfceaabeb+12tafaemmnen,\displaystyle\frac{1}{2}t_{c}f^{c}{}_{ab}e^{a}\wedge e^{b}+\frac{1}{2}t_{a}f^{a}{}_{mn}e^{m}\wedge e^{n}, (C.10)
demtm\displaystyle de^{m}t_{m} =\displaystyle= tmfmeaanen.\displaystyle t_{m}f^{m}{}_{an}e^{a}\wedge e^{n}. (C.11)

The first equation tells us that the field strength is given as

fαβa=faeαmmneβn.f^{a}_{\alpha\beta}\ =\ f^{a}{}_{mn}e^{m}_{\alpha}e^{n}_{\beta}. (C.12)

On the other hand, the second equation gives us the spin connection

ωα=mnfmeαaan.\omega_{\alpha}{}^{m}{}_{n}\ =\ -f^{m}{}_{an}e^{a}_{\alpha}. (C.13)

Then, the curvature

12Rαβdmnyαdyβ:=dωm+nωmkωkn\frac{1}{2}R_{\alpha\beta}{}^{m}{}_{n}dy^{\alpha}\wedge dy^{\beta}\ :=\ d\omega^{m}{}_{n}+\omega^{m}{}_{k}\wedge\omega^{k}{}_{n} (C.14)

can be also calculated explicitly. We obtain

Rαβ=mnfafmkleαkaneβl.R_{\alpha\beta}{}^{m}{}_{n}\ =\ -f^{a}{}_{kl}f^{m}{}_{an}e^{k}_{\alpha}e^{l}_{\beta}. (C.15)

Appendix D Equations of motion for the background flux

In this appendix, we show that the background gauge field A¯α\bar{A}_{\alpha} defined in section 2.3 satisfies the equations of motion with respect to the vielbein defined in Appendix C.

The equations of motion are

αF¯αβ+i[A¯α,F¯αβ]= 0,\nabla^{\alpha}\bar{F}_{\alpha\beta}+i[\bar{A}^{\alpha},\bar{F}_{\alpha\beta}]\ =\ 0, (D.1)

where

A¯α=eαaTa,F¯αβ=eαmeβnfaTamn.\bar{A}_{\alpha}\ =\ e^{a}_{\alpha}T_{a},\hskip 28.45274pt\bar{F}_{\alpha\beta}\ =\ e^{m}_{\alpha}e^{n}_{\beta}f^{a}{}_{mn}T_{a}. (D.2)

The covariant derivative is defined with respect to the spin connection as

αF¯αβ\displaystyle\nabla^{\alpha}\bar{F}_{\alpha\beta} =\displaystyle= eβn(mF¯mnωmlF¯lnmωmlF¯mln)\displaystyle e^{n}_{\beta}\left(\partial^{m}\bar{F}_{mn}-\omega^{ml}{}_{m}\bar{F}_{ln}-\omega^{ml}{}_{n}\bar{F}_{ml}\right) (D.3)
=\displaystyle= eβnemαeαa(flfbam+lnflfban)mlTb.\displaystyle e^{n}_{\beta}e^{m\alpha}e^{a}_{\alpha}\left(f^{l}{}_{am}f^{b}{}_{ln}+f^{l}{}_{an}f^{b}{}_{ml}\right)T_{b}.

The commutator can be written as

i[A¯α,F¯αβ]=eβnemαeαafbfcacTbmn.i[\bar{A}^{\alpha},\bar{F}_{\alpha\beta}]\ =\ -e^{n}_{\beta}e^{m\alpha}e^{a}_{\alpha}f^{b}{}_{ac}f^{c}{}_{mn}T_{b}. (D.4)

By using the Jaocbi identity and the fact that G/HG/H is symmetric, we find that the equations of motion (D.1) are automotaically satisfied.

Appendix E Explicit constructions of backgrounds in the SU(2)/U(1)\rm{SU}(2)/{\rm U}(1) coset space

In this appendix, we explicitly calculate eαme^{m}_{\alpha} and eαae^{a}_{\alpha} defined in Appendix C for the case G/H=S2G/H=S^{2}. We will see that eαme^{m}_{\alpha} is the standard zweibein which gives the round metric on S2S^{2}, and eαae^{a}_{\alpha} describes a monopole configuration on S2S^{2}.

Any element gg of SU(2)\rm SU(2) can be written as

g=[abba],|a|2+|b|2= 1.g\ =\ \left[\begin{array}[]{cc}a&-b^{*}\\ b&a^{*}\end{array}\right],\hskip 28.45274pt|a|^{2}+|b|^{2}\ =\ 1. (E.1)

The Maurer-Cartan 1-form g1dgg^{-1}dg is then given as

g1dg=σ+(bdaadb)+σ(adbbda)+σ3(ada+bdb),g^{-1}dg\ =\ \sigma_{+}(b^{*}da^{*}-a^{*}db^{*})+\sigma_{-}(adb-bda)+\sigma_{3}(a^{*}da+b^{*}db), (E.2)

where σ±\sigma_{\pm} and σ3\sigma_{3} are the Pauli matrices.

We choose a local section g(θ,φ)g(\theta,\varphi) by restricting a,ba,b to be

a=cosθ2,b=eiφsinθ2,a\ =\ \cos\frac{\theta}{2},\hskip 28.45274ptb\ =\ e^{i\varphi}\sin\frac{\theta}{2}, (E.3)

where θ,φ\theta,\varphi are the polar and the azimuthal angles of S2S^{2}, respectively. The pull-back of g1dgg^{-1}dg by this local section is then

g(θ,φ)1dg(θ,φ)=ie+t++iet+ie3t3,g(\theta,\varphi)^{-1}dg(\theta,\varphi)\ =\ ie^{+}t_{+}+ie^{-}t_{-}+ie^{3}t_{3}, (E.4)

where

t±:=σ±,t3:=12σ3,t_{\pm}\ :=\ \sigma_{\pm},\hskip 28.45274ptt_{3}\ :=\ \frac{1}{2}\sigma_{3}, (E.5)

and

e±:=±i2eiφ(dθisinθdφ),e3:=(1cosθ)dφ.e^{\pm}\ :=\ \pm\frac{i}{2}e^{\mp i\varphi}(d\theta\mp i\sin\theta d\varphi),\hskip 28.45274pte^{3}\ :=(1-\cos\theta)d\varphi. (E.6)

The metric hαβh_{\alpha\beta} obtained from e±e^{\pm} is therefore the round metric

ds2= 4e+e=dθ2+sin2θdφ2,ds^{2}\ =\ 4e^{+}e^{-}\ =\ d\theta^{2}+\sin^{2}\theta d\varphi^{2}, (E.7)

as expected. The field strength obtained from e3e^{3} is

de3=sinθdθdφ.de^{3}\ =\ \sin\theta\,d\theta d\varphi. (E.8)

The integral of this 2-form gives

12πS2sinθdθdφ= 2.\frac{1}{2\pi}\int_{S^{2}}\sin\theta\,d\theta d\varphi\ =\ 2. (E.9)

This shows that eα3e^{3}_{\alpha} describes a U(1)\rm U(1) monopole configuration with the monopole charge 22.

Note that the radius of S2S^{2} is set to be 11 in the above expressions. It is rather easy to recover the radius aa based on the dimensional analysis since aa is essentially the only dimensionful parameter in the Kaluza-Klein reduction of Yang-Mills theory. In higher dimensions, the coupling constant gYMg_{\rm YM} is dimensionful, but it is just an overall coefficient in the action.

In the following, we present an explicit description of the model of S2S^{2}=SU(2)/(2)/U(1)(1) compactification. We will employ different convensions from those in the main body of this paper, which might be more familiar to the readers.

The action of the six-dimensional space-time is

S=d6xg{1κ212g2Tr(FMNFMN)Λ},S=\int d^{6}x\sqrt{-g}\left\{\frac{1}{\kappa^{2}}{\cal R}-\frac{1}{2g^{2}}{\rm Tr}\left(F^{MN}F_{MN}\right)-\Lambda\right\}, (E.10)

where {\cal R} and Λ\Lambda are Ricci scalar and cosmological constant, respectively, and

FMN=MANNAMi[AM,AN]F_{MN}=\nabla_{M}A_{N}-\nabla_{N}A_{M}-i[A_{M},A_{N}] (E.11)

is the gauge field strength with the covariant derivative

MAN=MANΓLALMN\nabla_{M}A_{N}=\partial_{M}A_{N}-\Gamma^{L}{}_{MN}A_{L} (E.12)

and M,N=0,1,2,,4,5M,N=0,1,2,\cdots,4,5. The field strength is matrix valued as FMN=TaFMNaF_{MN}=T^{a}F^{a}_{MN}, where TaT^{a} is the generator matrix of the gauge group GYM=G_{\rm YM}=  SU(3)(3) with

Tr(TaTb)=12δab,[Ta,Tb]=ifabcTc.{\rm Tr}(T^{a}T^{b})=\frac{1}{2}\delta^{ab},\qquad[T^{a},T^{b}]=if^{abc}T^{c}. (E.13)

The metric is described as

ds2=dt2+i=1,2,3dxidxi+a2(dθ2+sin2θdφ2).ds^{2}=-dt^{2}+\sum_{i=1,2,3}dx^{i}dx^{i}+a^{2}\left(d\theta^{2}+\sin^{2}\theta d\varphi^{2}\right). (E.14)

The sphere of radius aa is described by two angular coordinates, θ\theta and φ\varphi, and non-compact spacetime is the Minkowski space. A vector on the sphere is described by two independent basis vactors, 𝐞θ{\bf e}_{\theta} and 𝐞φ{\bf e}_{\varphi}, which correspond to the two of three unit vectors of the polar coordinate system of three dimensional flat space.

Introduce a background gauge field on the sphere

𝐀¯=A¯θ𝐞θ+A¯φ𝐞φ,A¯θ=0,A¯φ=12Hcosθ1asinθ,{\bar{\bf A}}={\bar{A}}_{\theta}{\bf e}_{\theta}+{\bar{A}}_{\varphi}{\bf e}_{\varphi},\qquad{\bar{A}}_{\theta}=0,\qquad{\bar{A}}_{\varphi}=\frac{1}{2}H\frac{\cos\theta\mp 1}{a\sin\theta}, (E.15)

where negative sign is for 0<θ<π/20<\theta<\pi/2 and positive sign is for π/2<θ<π\pi/2<\theta<\pi in AφA_{\varphi}. Note that A¯4=aA¯θ{\bar{A}}_{4}=a{\bar{A}}_{\theta} and A¯5=asinθA¯φ{\bar{A}}_{5}=a\sin\theta{\bar{A}}_{\varphi}, since x4=θx_{4}=\theta and x5=φx_{5}=\varphi. This is a spherical slice of monopole configuration with unit magnetic charge at radius aa, and the generator of corresponding U(1)(1) gauge symmetry is H/2H/2. We can explicitly write

H=(n1000n2000n1n2).H=\left(\begin{array}[]{ccc}n_{1}&0&0\\ 0&n_{2}&0\\ 0&0&-n_{1}-n_{2}\end{array}\right). (E.16)

The off-diagonal components of the matrix-valued SU(3)(3) gauge field transform as matter fields under this U(1)(1) gauge symmetry with charges

((n1n2)/2(2n1+n2)/2(n1n2)/2(n1+2n2)/2(2n1+n2)/2(n1+2n2)/2),\left(\begin{array}[]{ccc}*&(n_{1}-n_{2})/2&(2n_{1}+n_{2})/2\\ -(n_{1}-n_{2})/2&*&(n_{1}+2n_{2})/2\\ -(2n_{1}+n_{2})/2&-(n_{1}+2n_{2})/2&*\end{array}\right), (E.17)

and the Dirac quantization of electric charge by a monopole indicates that twice of each charge should be an integer, which indicates

n1n2𝐙,3n1𝐙,3n2𝐙.n_{1}-n_{2}\in{\bf Z},\qquad 3n_{1}\in{\bf Z},\qquad 3n_{2}\in{\bf Z}. (E.18)

This background configuration is invariant under the symmetry transformation of S2S^{2}, or SU(2)\rm SU(2), up to corresponding gauge transformation generated by H/2H/2. Furthermore, we can make this background configuration a solution of field equations of the action, (E.10), namely SU(3)(3) Yang-Mills equation and Einstein equation by choosing Λ=1/κ2a2\Lambda=1/\kappa^{2}a^{2} and g2=Tr((H/2)2)κ4Λg^{2}={\rm Tr}((H/2)^{2})\kappa^{4}\Lambda [42].

The fluctuation around the background configuration of (E.15), δA4\delta A^{4} and δA5\delta A^{5}, can be described as scalar fields in low-energy four-dimensional effective theory. It is convenient to describe these field so that they are living on the tangent space of S2S^{2}. Introduce zweibein as

gμν=eμδmnmen,νg_{\mu\nu}=e_{\mu}{}^{m}\,\delta_{mn}\,e^{n}{}_{\nu}, (E.19)

where μ,ν=4,5\mu,\nu=4,5, m,n=4,5m,n=4,5 and in the matrix form gμν=a2diag(1,sin2θ)g_{\mu\nu}=a^{2}{\rm diag}(1,\sin^{2}\theta). Explicitly we specially introduce

en=νa(cosφsinφsinφcosφ)(100sinθ)e^{n}{}_{\nu}=a\left(\begin{array}[]{cc}\cos\varphi&-\sin\varphi\\ \sin\varphi&\cos\varphi\end{array}\right)\left(\begin{array}[]{cc}1&0\\ 0&\sin\theta\end{array}\right) (E.20)

which satisfies the above formula of definition. The fields on tangent space is defined as

VnenδνAν.V^{n}\equiv e^{n}{}_{\nu}\,\delta A^{\nu}. (E.21)

Furthermore, it is convenient to define a complex scalar field as

V±12(V4iV5),V=(V+).V_{\pm}\equiv\frac{1}{\sqrt{2}}\left(V^{4}\mp iV^{5}\right),\qquad V_{-}=(V_{+})^{\dagger}. (E.22)

We define a covariant derivative with background field as

D¯μV±μV±i[A¯μ,V±].{\bar{D}}_{\mu}V_{\pm}\equiv\nabla_{\mu}V_{\pm}-i[{\bar{A}}_{\mu},V_{\pm}]. (E.23)

The explicit form can be obtained as

D¯μV±=μV±iδμ5(cosθ1){±V±+[H2,V±]}{\bar{D}}_{\mu}V_{\pm}=\partial_{\mu}V_{\pm}-i\,\delta^{5}_{\mu}\,(\cos\theta-1)\left\{\pm V_{\pm}+\left[\frac{H}{2},V_{\pm}\right]\right\} (E.24)

for 0<θ<π/20<\theta<\pi/2. In case of π/2<θ<π\pi/2<\theta<\pi, the factor (cosθ1)(\cos\theta-1) in the second term should be replaced by (cosθ+1)(\cos\theta+1). Note that the field V±V_{\pm} is matrix valued as V±aTaV_{\pm}^{a}T^{a} and the second term can vanish depending on the choice of HH.

Appendix F Mode expansions on G/HG/H

In order to perform the Kaluza-Klein reduction on G/HG/H, we need to have a complete set of functions on G/HG/H by which any fields can be expanded. More precisely, we need to expand sections of suitable vector bundles on G/HG/H, not only functions, since there are fields with a local Lorentz indices belonging to a non-trivial representation of the gauge group. For S2S^{2}, the monopole harmonics [43] play such a role. In more general cases, it is known that there exists a useful set of functions on GG [40] which can be employed for our purpose, as we will review in this appendix.

Mode expansions on GG
It is rather easy to find examples of functions on GG. Let ρ:GGL(n,)\rho:G\to{\rm GL}(n,\mathbb{C}) be an nn-dimensional representation of GG. This assigns to each element gg of GG an n×nn\times n matrix ρ(g)\rho(g). Each matrix component ρ(g)IJ\rho(g)_{IJ} is therefore a function on GG. For each nn, or more appropriately, for each representation RR of GG, we can obtain many functions on GG in this manner. It can be shown that they are linearly independent.

The Peter-Weyl theorem tells us that any function on GG can be expanded by these functions. Explicitly, a function f(g)f(g) on GG can be written as

f(g)=RI,J=1dRc~IJRρR(g)IJ,f(g)\ =\ \sum_{R}\sum_{I,J=1}^{d_{R}}\tilde{c}^{R}_{IJ}\,\rho^{R}(g)_{IJ}, (F.1)

where the first sum is taken over all the representations of GG with the multiplicity one, dRd_{R} is the dimension of the representation RR, and ρR(g)\rho^{R}(g) is the matrix representing gGg\in G on the representation RR. For a later purpose, we employ an equivalent expansion

f(g)=RI,J=1dRcIJRρR(g1)IJ.f(g)\ =\ \sum_{R}\sum_{I,J=1}^{d_{R}}c^{R}_{IJ}\,\rho^{R}(g^{-1})_{IJ}. (F.2)

This is also valid since any function f(g)f(g) on GG can be written as f~(g1)\tilde{f}(g^{-1}) by using some f~(g)\tilde{f}(g).

Mode expansions of scalar functions on G/HG/H
We then study a complete set of functions on G/HG/H by using its extension to GG as follows. Let ϕ(y)\phi(y) be a function on a local coordinate patch UU of G/HG/H. By using a local section g(y)Gg(y)\in G, we can regard ϕ(y)\phi(y) as a function Φ(g(y))\Phi(g(y)) defined only on a subset of GG which is the embedding of UU. We can extend Φ(g(y))\Phi(g(y)) to the fiber direction by an action h(y)Hh(y)\in H as

Φ(g(y)h(y))=ϕ(y),h(y)H.\Phi(g(y)h(y))\ =\ \phi(y),\hskip 28.45274pth(y)\in H. (F.3)

In this manner, any function ϕ(y)\phi(y) on G/HG/H can be extended to a function Φ(g)\Phi(g) on GG. Thus a complete set of functions on G/HG/H can be obtained from a complete set of functions on GG by imposing the condition

Φ(g(y)h(y))=Φ(g(y)).{\Phi(g(y)h(y))\ =\ \Phi(g(y)).} (F.4)

Namely, these functions must take constant values along the fiber direction corresponding to HH-transformations.

Any function on GG can be uniquely expanded as (F.2). The representation RR of GG is decomposed into various irreducible representations of HH, which may contain unit representation 𝟏{\bf 1}. A complete set of scalar functions ϕ(y)\phi(y) on G/HG/H is then given by a complete set of constant functions along the fiber HH, which correspond to the unit representations constructed from all the representations RR of GG. An explicit form of the expansion is discussed as a special case of non-scalar functions (sections) which transform nontrivially as representation rr of HH.

Mode expansions of aμa_{\mu} and ϕm\phi_{m} on G/HG/H
In the Kaluza-Klein reduction of Yang-Mills theory, we would llike to expand aμa_{\mu} and ϕm\phi_{m}. Since they are sections of some vector bundles on G/HG/H and transforms nontrivially under HH, we need to modify the above expansion procedure as follows.

First, consider aμa_{\mu}. This belongs to the adjoint representation adYM\rm ad_{YM} of GYMG_{\rm YM}. Recall that, as explained in Appendix B, our principal GYMG_{\rm YM}-bundle is constructed from the principal HH-bundle GG, and the transition functions take values in a subgroup HH of GYMG_{\rm YM}. This means that we can consider aμa_{\mu} as a field defined on the principal HH-bundle GG. Since adYM\rm ad_{YM} is reducible for HH, we decompose aμa_{\mu} into components according to the irreducible decomposition of adYM\rm ad_{YM} with respect to HH. Each component belongs to an irreducible representation of HH, and forms a section of a vector bundle on G/HG/H. Note that this decomposition is compatible with the gauge symmetry preserved by the background flux F¯αβ\bar{F}_{\alpha\beta} since the preserved symmetry corresponds to a subgroup of GYMG_{\rm YM} which commutes with HH.

The case for ϕm\phi_{m} requires one more twist since it has the tangent space index mm, and transforms under the local Lorentz transformations. As explained under (2.21), the covariant derivative α\nabla_{\alpha} with respect to the metric on G/HG/H can be identified with the one with respect to the background gauge field in the representation RtR_{t} of HH, which is given by the commutation relation (C.7). Thus, the tangent space indices are regarded as indices of the representation RtR_{t} of HH. Then, ϕm\phi_{m} belongs to a product representation RtadYMR_{t}\otimes{\rm ad_{YM}} of HH. We decompose ϕm\phi_{m} into components according to the irreducible decomposition of RtadYMR_{t}\otimes{\rm ad_{YM}} with respect to HH. Each component again forms a section of a vector bundle on G/HG/H.

To summerize, the mode expansions of aμa_{\mu} and ϕm\phi_{m} can be performed if we know how to expand a section of a vector bundle on G/HG/H which belongs to an irreducible representation of HH. The mode expansions of aμa_{\mu} or ϕm\phi_{m} are given by a sum of various mode functions corresponding to each irreducible representations of HH.

Mode expansions of χi(y)\chi_{i}(y) in representation rr of HH
Now, we consider the expansion of χi(y)\chi_{i}(y) which belongs to a specific representation rr of HH. Recall a local section g(y)g(y) on G/HG/H can be extended to the fiber direction by the HH-transformation as

g(y)g(y)h(y).g(y)\to g(y)h(y).

Thus, any gGg\in G is written as a product of the local section g(y)g(y) and the HH-transformation h(y)h(y). This induces an HH-transformation on χi(y)\chi_{i}(y) as

χi(y)ρr(h(y)1)ijχj(y).\chi_{i}(y)\to\rho^{r}(h(y)^{-1})_{ij}\chi_{j}(y).

Thus, we can extend the function χi(y)\chi_{i}(y) on G/HG/H to a function 𝒳i(g){\cal X}_{i}(g) on GG by

𝒳i(g):=𝒳i(g(y)h(y))=ρr(h(y)1)ijχj(y),{\cal X}_{i}(g)\ :=\ {\cal X}_{i}(g(y)h(y))\ =\ \rho^{r}(h(y)^{-1})_{ij}\chi_{j}(y), (F.5)

where g=g(y)h(y)g=g(y)h(y). If rr is the trivial representation of HH, then this reduces to (F.3). By construction, 𝒳i(g){\cal X}_{i}(g) satisfies

𝒳i(gh)=ρr(h1)ij𝒳j(g).{\cal X}_{i}(gh)\ =\ \rho^{r}(h^{-1})_{ij}{\cal X}_{j}(g). (F.6)

Since each component 𝒳i(g){\cal X}_{i}(g) is a function on GG, it is expanded as (F.2). Then, the above condition (F.6) imposes the following restriction on the expansion coefficients cIJRc^{R}_{IJ}, namely an allowed set of functions on GG. For a representation RR of GG, the function ρR(g1)IJ\rho^{R}(g^{-1})_{IJ} on GG satisfies the transformation law

ρR((gh)1)IJ=ρR(h1g1)IJ=ρR(h1)IKρR(g1)KJ\rho^{R}((gh)^{-1})_{IJ}\ =\ \rho^{R}(h^{-1}g^{-1})_{IJ}\ =\ \rho^{R}(h^{-1})_{IK}\rho^{R}(g^{-1})_{KJ} (F.7)

under an action of hHh\in H. This representation RR of GG can be decomposed into irreducible representations r1rlr_{1}\oplus\cdots\oplus r_{l} of HH. For the basis according to this decomposition, the matrix ρR(h1)IK\rho^{R}(h^{-1})_{IK} takes a block-diagonal form. Let i1,,ili_{1},\cdots,i_{l} be indices corresponding to the representations r1,,rlr_{1},\cdots,r_{l}, respectively. Then, ρR(g1)i1K,,ρR(g1)ilK\rho^{R}(g^{-1})_{i_{1}K},\cdots,\rho^{R}(g^{-1})_{i_{l}K} transform separately as

ρR((gh)1)iaJ=ρra(h1)iajaρR(g1)jaJ,(a=1,,l)\rho^{R}((gh)^{-1})_{i_{a}J}\ =\ \rho^{r_{a}}(h^{-1})_{i_{a}j_{a}}\rho^{R}(g^{-1})_{j_{a}J},\hskip 28.45274pt(a=1,\cdots,l) (F.8)

under the HH-transformation. Let us first consider the case when the representation r1r_{1} is isomorphic to rr. Then a rectangular part ρR(g1)i1J\rho^{R}(g^{-1})_{i_{1}J} satisfies the condition (F.6). In general, the decomposition of RR contains a multiple of rr-representation of HH. Suppose that ra=rr_{a}=r for a=1,,ka=1,\cdots,k among (r1,rl).(r_{1},\cdots r_{l}). We denote the corresponding rectangular part by ρR,a(g1)iJ\rho^{R,a}(g^{-1})_{iJ} where ii is the index for rr. Then, the function 𝒳i(g){\cal X}_{i}(g) in the representation rr can be expanded in terms of these functions as

𝒳i(g)=rRJ=1dRa=1kcJR,aρR,a(g1)iJ.{\cal X}_{i}(g)\ =\ \sum_{r\subset R}\sum_{J=1}^{d_{R}}\sum_{a=1}^{k}c^{R,a}_{J}\,\rho^{R,a}(g^{-1})_{iJ}. (F.9)

The first sum is taken over all the representations RR of GG whose irreducible decomposition with respect to HH contains rr. Finally, the expansion of χi(y)\chi_{i}(y) is given as

χi(y)=rRJ=1dRa=1kcJR,afiJR,a(y),\chi_{i}(y)\ =\ \sum_{r\subset R}\sum_{J=1}^{d_{R}}\sum_{a=1}^{k}c^{R,a}_{J}\,f^{R,a}_{iJ}(y), (F.10)

where

fiJR,a(y):=ρR,a(g(y)1)iJ.f^{R,a}_{iJ}(y)\ :=\ \rho^{R,a}(g(y)^{-1})_{iJ}. (F.11)

These functions satisfy the same transformation law of (F.8) where ra=rr_{a}=r. In this paper, we sometimes suppress the index aa for notational simplicity.

Appendix G Laplacian and the mass formula on G/HG/H

In the Kaluza-Klein reduction, the mass of each mode is typically related to the eigenvalue of the Laplacian of the compactification manifold. The eigenvalues of the Laplacian on coset spaces were discussed in [51]. Interestingly, the mode functions we introduced in Appendix F turn out to be the eigenfunctions of the Laplacian on G/HG/H provided that the vielbein eαme^{m}_{\alpha} and the background gauge field A¯α\bar{A}_{\alpha} are given as in section 2 and Appendix C [36, 31].

First, we show that the action of the covariant derivative D¯α\bar{D}_{\alpha} on the mode functions fIJR(y)f^{R}_{IJ}(y) can be written in an algebraic form. Since fIJR(y)f^{R}_{IJ}(y) can be written as (ρR(g(y))1)IJ\left(\rho^{R}(g(y))^{-1}\right)_{IJ}, the exterior derivative is given as

dρR(g(y))1=ρR(g(y))1dρR(g(y))ρR(g(y))1,d\rho^{R}(g(y))^{-1}\ =\ -\rho^{R}(g(y))^{-1}d\rho^{R}(g(y))\cdot\rho^{R}(g(y))^{-1}, (G.1)

where the matrix indices are suppressed. The right-hand side contains the pull-back of the Maurer-Cartan 1-form in the representation RR. This can be expanded as

ρR(g(y))1dρR(g(y))=iemTmR+ieaTaR,\rho^{R}(g(y))^{-1}d\rho^{R}(g(y))\ =\ ie^{m}T^{R}_{m}+ie^{a}T^{R}_{a}, (G.2)

where TmR,TaRT^{R}_{m},T^{R}_{a} are the generators of 𝔤\mathfrak{g} in the representation RR. Inserting this expression into (G.1), we obtain

dρR(g(y))1+ieaTaRρR(g(y))1=iemTmRρR(g(y))1.d\rho^{R}(g(y))^{-1}+ie^{a}T^{R}_{a}\rho^{R}(g(y))^{-1}\ =\ -ie^{m}T^{R}_{m}\rho^{R}(g(y))^{-1}. (G.3)

Since the background gauge field is given by A¯=eαaTadyα\bar{A}=e^{a}_{\alpha}T_{a}dy^{\alpha}, this can be written as

D¯αfIJR(y)=ieαm(TmR)IKfKJR(y).\bar{D}_{\alpha}f^{R}_{IJ}(y)\ =\ -ie^{m}_{\alpha}(T^{R}_{m})_{IK}f^{R}_{KJ}(y). (G.4)

Thus, the group theoretic argument shows that the covariant derivative D¯α\bar{D}_{\alpha} on fIJR(y)f^{R}_{IJ}(y) can be simply written as a multiplication of ieαmTmR-ie^{m}_{\alpha}T^{R}_{m}. Note that the covariant derivative D¯α\bar{D}_{\alpha} originally contains the spin connection on G/HG/H, but as explained below (2.21), we can regard the local Lorentz indices as a charge of the gauge group HH on G/HG/H. Thus the spin connection term in D¯α\bar{D}_{\alpha} is absorbed into the gauge connection. The representation matrices of HH are thus given by the tensor product of the local Lorentz representation and the original HH charge.

In section 4, in particular in (4.17), we use this relation for G=SU(2)G={\rm SU}(2). The representations of SU(2){\rm SU}(2) are labeled by the half integers jj. The spin-jj representation is give by the matrix-valued function ρ(j)(g)mm\rho^{(j)}(g)_{mm^{\prime}} where jm,mj-j\leq m,m^{\prime}\leq j. We define the mode functions on G/HG/H by

fmmj(y):=ρ(j)(g(y)1)mm,f^{j}_{mm^{\prime}}(y)\ :=\ \rho^{(j)}(g(y)^{-1})_{mm^{\prime}}, (G.5)

where mm is constrained by the H=U(1)H=U(1) charge qq. The condition is discussed in (4.15).

The covariant derivative D¯+\bar{D}_{+} acting on fmmj(y)f^{j}_{mm^{\prime}}(y) can be rewritten as a multiplication of i(T+(j))mm-i\left(T^{(j)}_{+}\right)_{mm^{\prime}} on the function as (4.17). Note that the charge quantization discussed in Appendix E is automatically satisfied.

Next, we show that the mode functions fiJR(y)f^{R}_{iJ}(y) are eigenfunctions of the Laplacian hαβD¯αD¯βh^{\alpha\beta}\bar{D}_{\alpha}\bar{D}_{\beta}, where ii is the index for an irreducible representation rr of HH which is contained in RR [51]. Indeed,

hαβD¯αD¯βfiJR(y)\displaystyle-h^{\alpha\beta}\bar{D}_{\alpha}\bar{D}_{\beta}f^{R}_{iJ}(y) =\displaystyle= hαβD¯α(ieβn(TnR)iKfKJR(y))\displaystyle-h^{\alpha\beta}\bar{D}_{\alpha}\left(-ie^{n}_{\beta}(T^{R}_{n})_{iK}f^{R}_{KJ}(y)\right) (G.6)
=\displaystyle= ihαβeβn((TnR)iKαfKJR(y)+eαafm(TmR)iKanfKJR(y)+ieαa(TaRTnR)iKfKJR(y))\displaystyle ih^{\alpha\beta}e^{n}_{\beta}\left((T^{R}_{n})_{iK}\partial_{\alpha}f^{R}_{KJ}(y)+e^{a}_{\alpha}f^{m}{}_{an}(T^{R}_{m})_{iK}f^{R}_{KJ}(y)+ie^{a}_{\alpha}(T^{R}_{a}T^{R}_{n})_{iK}f^{R}_{KJ}(y)\right)
=\displaystyle= ihαβeβn(TmR)iKD¯αfKJR(y)+ieαa(ifmTmRan+[TaR,TnR])iKfKJR(y)\displaystyle ih^{\alpha\beta}e^{n}_{\beta}(T^{R}_{m})_{iK}\bar{D}_{\alpha}f^{R}_{KJ}(y)+ie^{a}_{\alpha}\left(-if^{m}{}_{an}T^{R}_{m}+[T^{R}_{a},T^{R}_{n}]\right)_{iK}f^{R}_{KJ}(y)
=\displaystyle= δmn(TmRTnR)iKfKJR(y)\displaystyle\delta^{mn}(T^{R}_{m}T^{R}_{n})_{iK}f^{R}_{KJ}(y)
=\displaystyle= (c2G(R)c2H(r))fiJR(y),\displaystyle\left(c^{G}_{2}(R)-c^{H}_{2}(r)\right)f^{R}_{iJ}(y),

where c2G(R)c_{2}^{G}(R) is the second Casimir invariant of the representation RR of GG.

This eigenvalue gives us the mass of each Kaluza-Klein mode of aμa_{\mu}. Recall that, for the mode expansion of aμa_{\mu} reviewed in Appendix F, we first need to decompose the adjoint representation adYM{\rm ad}_{\rm YM} of GYMG_{\rm YM} with respect to HH. Let rr be one of the irreducible representations of HH appearing in the decomposition of adYM\rm ad_{YM}. The components of aμa_{\mu} corresponding to rr are expanded in terms of fiJR(y)f^{R}_{iJ}(y) where RR is a representation of GG whose irreducibe decomposition with respect to HH contains rr. This vector mode has the mass given as

mv2=mR,r2:=c2G(R)c2H(r).m^{2}_{v}\ =\ m^{2}_{R,r}\ :=\ c^{G}_{2}(R)-c^{H}_{2}(r). (G.7)

Note that mR,r2m^{2}_{R,r} is always non-negative.

Next, we consider the masses of scalar modes obtained from ϕm\phi_{m}. Recall that the mass terms of the modes come from

Tr[12(D¯αϕβ)212ϕαRαβϕβiϕα[F¯αβ,ϕβ]].{\rm Tr}\left[\frac{1}{2}\left(\bar{D}_{\alpha}\phi_{\beta}\right)^{2}-\frac{1}{2}\phi^{\alpha}R_{\alpha\beta}\phi^{\beta}-i\phi^{\alpha}[\bar{F}_{\alpha\beta},\phi^{\beta}]\right]. (G.8)

The first term gives mR,r2m^{2}_{R,r}. The explicit expression of the curvature (C.15) implies that the Ricci tensor is given as

Rαβ=c2H(Rt)eαmeβntr(tmtn),R_{\alpha\beta}\ =\ -c_{2}^{H}(R_{t})e^{m}_{\alpha}e^{n}_{\beta}\,{\rm tr}(t_{m}t_{n}), (G.9)

where tmt_{m} belong to the representation used to define the coset space G/HG/H. Therefore, the curvature contribution to the mass is given as

12c2H(Rt)mTrnϕmϕn.\frac{1}{2}c_{2}^{H}(R_{t})^{m}{}_{n}\,{\rm Tr}\,\phi_{m}\phi^{n}. (G.10)

The flux contribution can be written as

ϕmA(TaRt)m(Ta)AnϕnBB,\phi_{mA}(T^{R_{t}}_{a})^{m}{}_{n}(T^{a})^{A}{}_{B}\phi^{nB}, (G.11)

where ϕm=ϕmATA\phi_{m}=\phi_{m}^{A}T_{A} is the expansion of ϕm\phi_{m} in terms of the generators TAT_{A} of GYMG_{\rm YM}, and (Ta)AB(T_{a})^{A}{}_{B} are the generators of 𝔥\mathfrak{h} represented on adYM{\rm ad}_{\rm YM} which is reducible with respect to 𝔥\mathfrak{h}. Like in the case of the angular momentum in quantum mechanics, this can be written as

12ϕmA(c2H(RtadYM)mAnBc2H(Rt)mδBAnc2H(adYM)AδnmB)ϕnB.\frac{1}{2}\phi_{mA}\left(c_{2}^{H}(R_{t}\otimes{\rm ad}_{\rm YM})^{mA}{}_{nB}-c_{2}^{H}(R_{t})^{m}{}_{n}\delta^{A}_{B}-c_{2}^{H}({\rm ad}_{\rm YM})^{A}{}_{B}\delta^{m}_{n}\right)\phi^{nB}. (G.12)

The second term cancels the curvature contribution.

In order to determine the mass explicitly, we consider the decomposition of ϕm\phi_{m} in more detail. This belongs to RtadYMR_{t}\otimes{\rm ad_{YM}}. First, we decompose adYM\rm ad_{YM} and pick up one irreducible representation r~\tilde{r} of HH. The corresponding components of ϕm\phi_{m} belong to Rtr~R_{t}\otimes\tilde{r}. We further decompose Rtr~R_{t}\otimes\tilde{r} and pick up rr. These components are expanded in terms of fiJR(y)f^{R}_{iJ}(y) where RR contains rr. Their masses are therefore given as

ms2=mR,r2+c2H(r)c2H(r~)=c2G(R)c2H(r~).m_{s}^{2}\ =\ m^{2}_{R,r}+c_{2}^{H}(r)-c_{2}^{H}(\tilde{r})\ =\ c_{2}^{G}(R)-c_{2}^{H}(\tilde{r}). (G.13)

The important difference of this mass formula from mv2m_{v}^{2} is that the second term in the right-hand side is the Casimir invariant for r~\tilde{r}, not for rr. For example, even if r~\tilde{r} is non-trivial, Rtr~R_{t}\otimes\tilde{r} may contain a singlet component. In this case, we can choose R=𝟏R={\bf 1}, so that m2m^{2} is negative.

Indeed, this happens for the symmetric Higgs fields. As defined in section 3, a symmetric Higgs field consists of those components of ϕm\phi_{m} which is singlet with respect to HH. This means r=𝟏r={\bf 1}. Therefore, the mode expansion of the symmetric Higgs field do contain the contribution from R=𝟏R={\bf 1}. This is nothing but the constant mode, and the mass is given as

m2=c2H(r~).m^{2}\ =\ -c_{2}^{H}(\tilde{r}). (G.14)

As long as r~\tilde{r} is non-trivial, this mode has a tachyonic mass term which allows the symmetric Higgs field to acquire a non-zero vacuum expectation value. This fact justifies their name.

Appendix H Symmetric Higgs fields are symmetric

In this Appendix, we show that the symmetric Higgs fields defined in section 3 are symmetric fields in the sense of [39], as the name suggests.

First, we introduce the notion of GG-invariant fields. A field ϕm\phi_{m} on G/HG/H is said to be GG-invariant if this satisfies

g0ϕm(y)=ϕm(y),g0G,g_{0}\cdot\phi_{m}(y)\ =\ \phi_{m}(y),\hskip 28.45274ptg_{0}\in G, (H.1)

where the action of g0g_{0} is defined as

g0ϕm(y):=Φm(g0g(y)),g_{0}\cdot\phi_{m}(y)\ :=\ \Phi_{m}(g_{0}g(y)), (H.2)

where Φm(g)\Phi_{m}(g) are a set of functions on GG which extend ϕm(y)\phi_{m}(y) on G/HG/H to the fiber directions HH, as explained in Appendix F. Recall that g0g(y)g_{0}g(y) can be written as g(y)h(y,g0)g(y^{\prime})h(y,g_{0}), where yyy\to y^{\prime} is an isometry of G/HG/H induced by g0g_{0} and h(y,g0)Hh(y,g_{0})\in H is an HH-transformation. Then, the condition (H.1) can be written as

Λmn(y)U(y)ϕn(y)U(y)=ϕm(y),\Lambda_{mn}(y)U(y)\phi_{n}(y^{\prime})U(y)^{\dagger}\ =\ \phi_{m}(y), (H.3)

where Λmn(y)\Lambda_{mn}(y) and U(y)U(y) are the local Lorentz transformation and the gauge transformation induced by h(y,g0)h(y,g_{0}). This shows that a GG-invariant field is symmetric.

Recall that a symmetric Higgs field ϕm\phi_{m} is defined to satisfy D¯αϕn=0\bar{D}_{\alpha}\phi_{n}=0 and αϕn=0\partial_{\alpha}\phi_{n}=0. These imply that ϕm\phi_{m} is yαy^{\alpha}-independent and HH-invariant. Then, we find

Λmn(y)U(y)ϕn(y)U(y)=Λmn(y)U(y)ϕn(y)U(y)=ϕm(y).\Lambda_{mn}(y)U(y)\phi_{n}(y^{\prime})U(y)^{\dagger}\ =\ \Lambda_{mn}(y)U(y)\phi_{n}(y)U(y)^{\dagger}\ =\ \phi_{m}(y). (H.4)

Therefore, ϕm\phi_{m} is GG-invariant. This then implies that it is symmetric.

References

  • [1] S. R. Coleman and E. J. Weinberg, “Radiative Corrections as the Origin of Spontaneous Symmetry Breaking,” Phys. Rev. D 7 (1973), 1888-1910.
  • [2] L. E. Ibanez and G. G. Ross, “SU(2)L×(2)_{L}\timesU(1)(1) Symmetry Breaking as a Radiative Effect of Supersymmetry Breaking in Guts,” Phys. Lett. B 110 (1982), 215-220.
  • [3] K. Inoue, A. Kakuto, H. Komatsu and S. Takeshita, “Aspects of Grand Unified Models with Softly Broken Supersymmetry,” Prog. Theor. Phys. 68 (1982), 927 [erratum: Prog. Theor. Phys. 70 (1983), 330].
  • [4] W. A. Bardeen, “On naturalness in the standard model,” FERMILAB-CONF-95-391-T.
  • [5] S. Iso, N. Okada and Y. Orikasa, “The minimal B-L model naturally realized at TeV scale,” Phys. Rev. D 80 (2009), 115007 [arXiv:0909.0128 [hep-ph]].
  • [6] S. Iso and Y. Orikasa, “TeV Scale B-L model with a flat Higgs potential at the Planck scale: In view of the hierarchy problem,” PTEP 2013 (2013), 023B08 [arXiv:1210.2848 [hep-ph]].
  • [7] I. Antoniadis, K. Benakli and M. Quiros, “Radiative symmetry breaking in brane models,” Nucl. Phys. B 583 (2000), 35-48 [arXiv:hep-ph/0004091 [hep-ph]].
  • [8] N. S. Manton, “A New Six-Dimensional Approach to the Weinberg-Salam Model,” Nucl. Phys. B 158, 141-153 (1979).
  • [9] D. B. Fairlie, “Two Consistent Calculations of the Weinberg Angle,” J. Phys. G 5 (1979), L55.
  • [10] D. B. Fairlie, “Higgs’ Fields and the Determination of the Weinberg Angle,” Phys. Lett. B 82 (1979), 97-100.
  • [11] Y. Hosotani, “Dynamical Mass Generation by Compact Extra Dimensions,” Phys. Lett. B 126 (1983), 309-313.
  • [12] Y. Hosotani, “Dynamical Gauge Symmetry Breaking as the Casimir Effect,” Phys. Lett. B 129 (1983), 193-197.
  • [13] Y. Hosotani, “Dynamics of Nonintegrable Phases and Gauge Symmetry Breaking,” Annals Phys. 190 (1989), 233.
  • [14] I. Antoniadis, “A Possible new dimension at a few TeV,” Phys. Lett. B 246 (1990), 377-384
  • [15] I. Antoniadis, N. Arkani-Hamed, S. Dimopoulos and G. R. Dvali, “New dimensions at a millimeter to a Fermi and superstrings at a TeV,” Phys. Lett. B 436 (1998), 257-263 [arXiv:hep-ph/9804398 [hep-ph]].
  • [16] H. Hatanaka, T. Inami and C. S. Lim, “The Gauge hierarchy problem and higher dimensional gauge theories,” Mod. Phys. Lett. A 13 (1998), 2601-2612. [arXiv:hep-th/9805067 [hep-th]].
  • [17] N. Arkani-Hamed, S. Dimopoulos and G. R. Dvali, “Phenomenology, astrophysics and cosmology of theories with submillimeter dimensions and TeV scale quantum gravity,” Phys. Rev. D 59 (1999), 086004 [arXiv:hep-ph/9807344 [hep-ph]].
  • [18] L. Randall and R. Sundrum, “A Large mass hierarchy from a small extra dimension,” Phys. Rev. Lett. 83 (1999), 3370-3373 [arXiv:hep-ph/9905221 [hep-ph]].
  • [19] L. Randall and R. Sundrum, “An Alternative to compactification,” Phys. Rev. Lett. 83 (1999), 4690-4693 [arXiv:hep-th/9906064 [hep-th]].
  • [20] G. R. Dvali, S. Randjbar-Daemi and R. Tabbash, “The Origin of spontaneous symmetry breaking in theories with large extra dimensions,” Phys. Rev. D 65, 064021 (2002) [arXiv:hep-ph/0102307 [hep-ph]].
  • [21] N. Arkani-Hamed, A. G. Cohen and H. Georgi, “Electroweak symmetry breaking from dimensional deconstruction,” Phys. Lett. B 513 (2001), 232-240 [arXiv:hep-ph/0105239 [hep-ph]].
  • [22] C. S. Lim, N. Maru and K. Hasegawa, “Six Dimensional Gauge-Higgs Unification with an Extra Space S**2 and the Hierarchy Problem,” J. Phys. Soc. Jap. 77 (2008), 074101 [arXiv:hep-th/0605180 [hep-th]].
  • [23] N. Kitazawa and S. Kobayashi, “Spontaneous Gauge Symmetry Breaking in a Non-Supersymmetric D-brane Model,” Phys. Lett. B 720 (2013), 373-378 [arXiv:1211.1777 [hep-th]].
  • [24] S. Iso and N. Kitazawa, “Revolving D-branes and Spontaneous Gauge Symmetry Breaking,” PTEP 2015 (2015) no.12, 123B01 [arXiv:1507.04834 [hep-ph]].
  • [25] D. Kapetanakis and G. Zoupanos, “Coset space dimensional reduction of gauge theories,” Phys. Rept. 219, 4-76 (1992).
  • [26] D. Andriot, A. Cornell, A. Deandrea, F. Dogliotti and D. Tsimpis, “A new mechanism for symmetry breaking from nilmanifolds,” JHEP 05 (2020), 122 [arXiv:2002.11128 [hep-th]].
  • [27] S. Randjbar-Daemi, A. Salam and J. A. Strathdee, “Instability of Higher Dimensional Yang-Mills Systems,” Phys. Lett. B 124, 345 (1983) [erratum: Phys. Lett. B 144, 455 (1984)].
  • [28] S. Randjbar-Daemi, A. Salam and J. A. Strathdee, “Instanton Induced Compactification and Fermion Chirality,” Phys. Lett. B 132, 56-60 (1983).
  • [29] S. Randjbar-Daemi, A. Salam and J. A. Strathdee, “Spontaneous Compactification in Six-Dimensional Einstein-Maxwell Theory,” Nucl. Phys. B 214, 491-512 (1983).
  • [30] S. Randjbar-Daemi, A. Salam and J. A. Strathdee, “Stability of Instanton Induced Compactification in Eight-dimensions,” Nucl. Phys. B 242, 447-472 (1984).
  • [31] A. N. Schellekens, “Stability of Higher Dimensional Einstein Yang-Mills Theories on Symmetric Spaces,” Nucl. Phys. B 248, 706-726 (1984).
  • [32] A. N. Schellekens, “COMPACTIFICATION OF EINSTEIN YANG-MILLS THEORIES ON CP(n),” Nucl. Phys. B 262, 661-688 (1985).
  • [33] M. R. Douglas and S. Kachru, “Flux compactification,” Rev. Mod. Phys. 79 (2007), 733-796 [arXiv:hep-th/0610102 [hep-th]].
  • [34] R. Blumenhagen, B. Kors, D. Lust and S. Stieberger, “Four-dimensional String Compactifications with D-Branes, Orientifolds and Fluxes,” Phys. Rept. 445 (2007), 1-193 [arXiv:hep-th/0610327 [hep-th]].
  • [35] M. Grana, “Flux compactifications in string theory: A Comprehensive review,” Phys. Rept. 423 (2006), 91-158 [arXiv:hep-th/0509003 [hep-th]].
  • [36] L. Palla, “On Dimensional Reduction of Gauge Theories: Symmetric Fields and Harmonic Expansion,” Z. Phys. C 24, 195 (1984).
  • [37] N. G. Kozimirov, V. A. Kuzmin and I. I. Tkachev, “ON SYMMETRY BREAKING AT COSET SPACE DIMENSIONAL REDUCTION,” Phys. Lett. B 215, 84-86 (1988).
  • [38] H. Itoyama, Y. Koga and S. Nakajima, “Target Space Duality of Non-Supersymmetric String Theory,” [arXiv:2110.09762 [hep-th]].
  • [39] P. Forgacs and N. S. Manton, “Space-Time Symmetries in Gauge Theories,” Commun. Math. Phys. 72, 15 (1980).
  • [40] A. Salam and J. A. Strathdee, “On Kaluza-Klein Theory,” Annals Phys. 141, 316-352 (1982).
  • [41] P. Forgacs, Z. Horvath and L. Palla, “SPONTANEOUS COMPACTIFICATION TO NONSYMMETRIC SPACES,” Z. Phys. C 30, 261 (1986).
  • [42] S. Randjbar-Daemi and R. Percacci, “Spontaneous Compactification of a (4+dd)-dimensional Kaluza-Klein Theory Into M(4) X G/HG/H for Arbitrary GG and HH,” Phys. Lett. B 117, 41 (1982).
  • [43] T. T. Wu and C. N. Yang, “Dirac Monopole Without Strings: Monopole Harmonics,” Nucl. Phys. B 107, 365 (1976).
  • [44] S. Weinberg, “The Quantum theory of fields. Vol. 1: Foundations,” Cambridge University Press.
  • [45] T. Eguchi, P. B. Gilkey and A. J. Hanson, “Gravitation, Gauge Theories and Differential Geometry,” Phys. Rept. 66, 213 (1980).
  • [46] J. P. Harnad, S. Shnider and J. Tafel, “Group Actions on Principal Bundles and Dimensional Reduction,” Lett. Math. Phys. 4, 107-113 (1980).
  • [47] G. Chapline and N. S. Manton, “The Geometrical Significance of Certain Higgs Potentials: An Approach to Grand Unification,” Nucl. Phys. B 184, 391-405 (1981).
  • [48] A. Adams, J. Polchinski and E. Silverstein, “Don’t panic! Closed string tachyons in ALE space-times,” JHEP 10, 029 (2001) [arXiv:hep-th/0108075 [hep-th]].
  • [49] T. Suyama, “Closed string tachyons in nonsupersymmetric heterotic theories,” JHEP 08, 037 (2001) [arXiv:hep-th/0106079 [hep-th]].
  • [50] S. Kobayashi, K. Nomizu, “Foundations of Differential Geometry, volume 1,” John Wiley & Sons, Inc.
  • [51] K. Pilch and A. N. Schellekens, “Formulae for the Eigenvalues of the Laplacian on Tensor Harmonics on Symmetric Coset Spaces,” J. Math. Phys. 25, 3455 (1984).