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

institutetext: Physics Division, National Center for Theoretical Sciences, National Tsing-Hua University,
Hsinchu 30013, Taiwan

Free energy and defect CC-theorem in free fermion

Yoshiki Sato
Abstract

We describe a pp-dimensional conformal defect of a free Dirac fermion on a dd-dimensional flat space as boundary conditions on a conformally equivalent space p+1×𝕊dp1\mathbb{H}^{p+1}\times\mathbb{S}^{d-p-1}. We classify allowed boundary conditions and find that the Dirichlet type of boundary conditions always exists while the Neumann type of boundary condition exists only for a two-codimensional defect. For the two-codimensional defect, a double trace deformation triggers a renormalization group flow from the Neumann boundary condition to the Dirichlet boundary condition, and the free energy at UV fixed point is always larger than that at IR fixed point. This provides us with further support of a conjectured CC-theorem in DCFT.

1 Introduction

A renormalization group (RG) is a fundamental concept in quantum field theory (QFT). By a deformation by a relevant operator 𝒪\mathcal{O},

ICFT+λddx𝒪,\displaystyle I_{\text{CFT}}+\lambda\int\!\mathrm{d}^{d}x\,\mathcal{O}\,, (1)

a conformal field theory (CFT) at a UV fixed point flows to a CFT at an IR fixed point. The RG flow is not reversible, and this irreversibility implies the existence of a monotonic function, known as a CC-function. A CC-theorem which states the existence of the CC-function is pioneered by Zamolodchikov Zamolodchikov:1986gt in two-dimensional QFTs and is extended to higher dimensions Cardy:1988cwa ; Komargodski:2011vj ; Jafferis:2011zi ; Klebanov:2011gs ; Myers:2010xs ; Myers:2010tj . Combining aa-theorem Cardy:1988cwa ; Komargodski:2011vj which is a CC-theorem in d=4d=4 and FF-theorem Jafferis:2011zi ; Klebanov:2011gs which is a CC-theorem in d=3d=3, a generalised FF-theorem is conjectured Giombi:2014xxa : A free energy on a sphere 𝕊d\mathbb{S}^{d},

F~=sin(πd2)logZ[𝕊d],\displaystyle\tilde{F}=\sin\left(\frac{\pi d}{2}\right)\log Z[\mathbb{S}^{d}]\,, (2)

is a CC-function and satisfies the monotonic relation

F~UVF~IR.\displaystyle\tilde{F}_{\text{UV}}\geq\tilde{F}_{\text{IR}}\,. (3)

The generalised FF-theorem states monotonicity of an anomaly coefficient in the free energy for even dd, while monotonicity of a finite part in the free energy for odd dd. The CC-theorem can also be proved by using an information theoretical method for d4d\leq 4 Casini:2004bw ; Casini:2012ei ; Casini:2017vbe .

For QFTs with pp-dimensional defect, it is possible to consider an RG flow triggered by a relevant operator localising on the defect,

IDCFT+λ^dpx𝒪^.\displaystyle I_{\text{DCFT}}+\hat{\lambda}\int\!\mathrm{d}^{p}x\,\hat{\mathcal{O}}\,. (4)

In Kobayashi:2018lil , we conjectured that the defect free energy on a sphere, which is an increment of the free energy due to the defect,

log𝒟(p)=logZDCFT[𝕊d]logZCFT[𝕊d],\displaystyle\log\langle\mathcal{D}^{(p)}\rangle=\log Z^{\text{DCFT}}[\mathbb{S}^{d}]-\log Z^{\text{CFT}}[\mathbb{S}^{d}]\,, (5)

decreases under the defect RG flow. More precisely, the universal part of the defect free energy,

D~=sin(πp2)log|𝒟(p)|\displaystyle\tilde{D}=\sin\left(\frac{\pi p}{2}\right)\log|\langle\mathcal{D}^{(p)}\rangle| (6)

is a CC-function, and it decreases under the defect RG flow,

D~UVD~IR.\displaystyle\tilde{D}_{\text{UV}}\geq\tilde{D}_{\text{IR}}\,. (7)

For BCFTs with p=d1p=d-1, a slight modification is needed since BCFTs is defined on a hemisphere 𝕊d\mathbb{HS}^{d} (a half space of the original CFTs), and the boundary free energy is introduced as111It is pointed out that the boundary free energy does not decrease monotonically under the bulk RG flow Green:2007wr ; Sato:2020upl .

log𝒟(p)=logZBCFT[𝕊d]12logZCFT[𝕊d].\displaystyle\log\langle\mathcal{D}^{(p)}\rangle=\log Z^{\text{BCFT}}[\mathbb{HS}^{d}]-\frac{1}{2}\log Z^{\text{CFT}}[\mathbb{S}^{d}]\,. (8)

For BCFT, our conjecture reproduces a proved CC-theorem in BCFT2, known as gg-theorem Affleck:1991tk ; Friedan:2003yc ; Casini:2016fgb , and a CC-theorem in higher dimensional BCFT Jensen:2015swa ; Nozaki:2012qd ; Gaiotto:2014gha ; Wang:2021mdq . In particular, the CC-theorem in BCFT3 is proved in Jensen:2015swa by extending Komargodski:2011vj and in Casini:2018nym by extending Casini:2016fgb . In the context of holography, a CC-theorem in BCFT is investigated in Yamaguchi:2002pa ; Takayanagi:2011zk ; Fujita:2011fp ; Estes:2014hka ; Miao:2017gyt .222In Estes:2014hka , it is argued that the defect entropy, which is an increment of the entanglement entropy of the spherical region where the defect sits at the centre due to the presence of the defect, is a candidate of a CC-function in DCFT. The defect entropy is not a CC-function for p<d1p<d-1 as pointed out Kobayashi:2018lil . See also related works Kumar:2016jxy ; Kumar:2017vjv . For DCFT with p<d1p<d-1, our conjecture however has passed only several checks in field theory Jensen:2015swa ; Beccaria:2017rbe ; Kobayashi:2018lil ; Jensen:2018rxu ; Wang:2020xkc ; Wang:2021mdq and holography Kumar:2016jxy ; Kumar:2017vjv ; Rodgers:2018mvq .

Recently, we provide further evidence of our conjecture in a simple model, a conformally coupled scalar field Nishioka:2021uef (See also related works Rodriguez-Gomez:2017kxf ; Rodriguez-Gomez:2017aca ; Gustavsson:2019zwm ). Instead of putting the conformally coupled scalar field on the sphere, we put it on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}, where q=dpq=d-p is a codimension of the defect, and impose boundary conditions at the boundary of p+1\mathbb{H}^{p+1}. The idea of mapping BCFTs on a flat space d\mathbb{R}^{d} to d\mathbb{H}^{d} has appeared in Herzog:2019bom ; Giombi:2020rmc , and the similar idea for DCFTs has appeared in Kapustin:2005py ; Chester:2015wao ; Rodriguez-Gomez:2017kxf . It enables us to classify allowed boundary conditions which are consistent with a recent classification Lauria:2020emq . The Dirichlet type boundary condition is always allowed while the Neumann type boundary condition is allowed only in q=1,2,3,4q=1,2,3,4. An RG flow from the Neumann boundary condition to the Dirichlet boundary condition realised by a double trace deformation as is familiar with the AdS/CFT setup Witten:2001ua ; Berkooz:2002ug ; Gubser:2002zh ; Gubser:2002vv ; Hartman:2006dy ; Diaz:2007an ; Giombi:2013yva , and the defect free energy with the Neumann boundary condition is always larger than that of the Dirichlet boundary condition.

In this paper, we extend our analysis Nishioka:2021uef to a free fermion to provide a further check of our conjecture. To compare with a scalar field, a defect free energy of a free fermion in higher dimensions has been studied only in BCFT Rodriguez-Gomez:2017aca ; Herzog:2019bom and DCFT with a two-codimensional defect in the context of entanglement entropy Klebanov:2011uf ; Beccaria:2017dmw . In particular, the existence of a nontrivial boundary condition is unclear in a free fermion since a unique boundary condition is allowed in d\mathbb{H}^{d} Dowker:1995sw ; Rodriguez-Gomez:2017aca .

The organisation of this paper is as follows. In the next section, we summarise coordinate systems and Weyl transformations among the coordinate systems. Furthermore, we classify boundary conditions of a massive fermion on d\mathbb{H}^{d} and a massless fermion on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}. In particular, for a massless fermion on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}, we show that a boundary condition of a Dirichlet type always exists while a boundary condition of a Neumann type exists only in q=2q=2. In section 3, we compute free energies of a massless fermion on 𝕊d\mathbb{S}^{d} and a hemisphere 𝕊d\mathbb{HS}^{d} using a zeta-function regularisation for a warm-up of the next section. In section 4, we compute free energies of a massive fermion on d\mathbb{H}^{d} and free energies of a massless fermion p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} with Dirichlet boundary condition by using a zeta-function regularisation. In section 5, we obtain free energies with Neumann boundary condition by analytical continuation and confirm the validity of our conjecture. The final section is devoted to discussion. Appendix A contains the list of anomaly parts and finite parts of free energies on 𝕊d\mathbb{S}^{d}, 𝕊d\mathbb{HS}^{d}, d\mathbb{H}^{d} and p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} with Dirichlet boundary condition, and appendix B is a technical details of a computation.

2 Classification of boundary conditions

2.1 Coordinate system

In this section, we summarise coordinate systems for a sphere 𝕊d\mathbb{S}^{d}, a hemisphere 𝕊d\mathbb{HS}^{d}, a hyperbolic space d\mathbb{H}^{d} and p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} and conformal maps among a flat space d\mathbb{R}^{d} and them.

Let us consider DCFTd on d\mathbb{R}^{d},

ds2=dxa2+dyi2,(a=1,,p,i=p+1,,d)\displaystyle\mathrm{d}s^{2}=\mathrm{d}x_{a}^{2}+\mathrm{d}y_{i}^{2}\,,\qquad(a=1,\cdots,p,i=p+1,\cdots,d) (9)

where a pp-dimensional defect sits at yi=0y_{i}=0. For later convenience, we introduce a codimension of the defect,

q=dp.\displaystyle q=d-p\,. (10)

By using the polar coordinate for the yiy_{i}-coordinate,

dyi2=dz2+z2ds𝕊q12,\displaystyle\mathrm{d}y_{i}^{2}=\mathrm{d}z^{2}+z^{2}\mathrm{d}s_{\mathbb{S}^{q-1}}^{2}\,, (11)

the metric of the flat space becomes

ds2=dxa2+dz2+z2ds𝕊q12=z2(dxa2+dz2z2+ds𝕊q12).\displaystyle\begin{aligned} \mathrm{d}s^{2}&=\mathrm{d}x_{a}^{2}+\mathrm{d}z^{2}+z^{2}\mathrm{d}s_{\mathbb{S}^{q-1}}^{2}\\ &=z^{2}\left(\frac{\mathrm{d}x_{a}^{2}+\mathrm{d}z^{2}}{z^{2}}+\mathrm{d}s_{\mathbb{S}^{q-1}}^{2}\right)\,.\end{aligned} (12)

After a Weyl rescaling, the metric (12) reduces to a geometry p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} with radius RR,

ds2=R2(dxa2+dz2z2+ds𝕊q12).\displaystyle\mathrm{d}s^{2}=R^{2}\left(\frac{\mathrm{d}x_{a}^{2}+\mathrm{d}z^{2}}{z^{2}}+\mathrm{d}s_{\mathbb{S}^{q-1}}^{2}\right)\,. (13)

Now the defect sits at the boundary of p+1\mathbb{H}^{p+1}. We can also use the global coordinate for p+1\mathbb{H}^{p+1},

ds2=R2(dρ2+sinh2ρds𝕊p2+ds𝕊q12),\displaystyle\mathrm{d}s^{2}=R^{2}\left(\mathrm{d}\rho^{2}+\sinh^{2}\rho\,\mathrm{d}s_{\mathbb{S}^{p}}^{2}+\,\mathrm{d}s_{\mathbb{S}^{q-1}}^{2}\right)\,, (14)

where the defect sits at ρ=\rho=\infty. Introducing a new variable φ\varphi, tanφ=sinhρ\tan\varphi=\sinh\rho, the metric (14) becomes

ds2=R2cos2φ(dφ2+sin2φds𝕊p2+cos2φds𝕊q12).\displaystyle\mathrm{d}s^{2}=\frac{R^{2}}{\cos^{2}\varphi}\left(\mathrm{d}\varphi^{2}+\sin^{2}\varphi\,\mathrm{d}s_{\mathbb{S}^{p}}^{2}+\cos^{2}\varphi\,\mathrm{d}s_{\mathbb{S}^{q-1}}^{2}\right)\,. (15)

After a Weyl rescaling, the metric (15) can be mapped to the sphere metric with radius RR

ds2=R2(dφ2+sin2φds𝕊p2+cos2φds𝕊q12),\displaystyle\mathrm{d}s^{2}=R^{2}\left(\mathrm{d}\varphi^{2}+\sin^{2}\varphi\,\mathrm{d}s_{\mathbb{S}^{p}}^{2}+\cos^{2}\varphi\,\mathrm{d}s_{\mathbb{S}^{q-1}}^{2}\right)\,, (16)

where 0φ<π0\leq\varphi<\pi and the defect sits at φ=π/2\varphi=\pi/2. The hemisphere 𝕊d\mathbb{HS}^{d} has the same metric of the sphere (16). A different point is that the range of φ\varphi is 0φπ/20\leq\varphi\leq\pi/2 and the boundary sits at φ=π/2\varphi=\pi/2.

2.2 Boundary condition of fermion on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}

In this section, we classify allowed boundary conditions of a massless fermion on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}. A free fermion on a curved background which is conformally equivalent to d\mathbb{R}^{d} is studied in e.g. Camporesi:1992tm ; Camporesi:1995fb ; Lewkowycz:2012qr ; Herzog:2019bom ; Klebanov:2011uf . See Chester:2015wao for the notation of a fermion on 2×𝕊2\mathbb{H}^{2}\times\mathbb{S}^{2}. The action of a massive Dirac fermion is given by333The massive fermion is not conformal. However, we introduce a mass term for later convenience.

I=ddxg(iψΓaaψ+Mψψ),\displaystyle I=\int\!\mathrm{d}^{d}x\sqrt{g}\,\left(\textrm{i}\,\psi^{\dagger}\,\Gamma^{a}\nabla_{a}\psi+M\psi^{\dagger}\psi\right)\,, (17)

where we assume M0M\geq 0. The rank of the gamma matrix for d2d\geq 2 is

rd=2d2,\displaystyle r_{d}=2^{\lfloor\frac{d}{2}\rfloor}\,, (18)

and the gamma matrix satisfies the anti-commutation relation,

{Γa,Γb}=2δab𝟙.\displaystyle\{\Gamma^{a},\Gamma^{b}\}=2\delta^{ab}\mathds{1}\,. (19)

The covariant derivative a\nabla_{a} and the spin connection ωμbc\omega_{\mu bc} are defined as

a=eaμμ,μ=μ+12σbcωμbc,σab=14[Γa,Γb],ωμbc=ebν(μecνΓνμαecα)\displaystyle\begin{aligned} \nabla_{a}&=e_{a}^{\mu}\nabla_{\mu}\,,\qquad\nabla_{\mu}=\partial_{\mu}+\frac{1}{2}\sigma^{bc}\omega_{\mu bc}\,,\\ \sigma^{ab}&=\frac{1}{4}[\Gamma^{a},\Gamma^{b}]\,,\qquad\omega_{\mu bc}=e_{b}^{\nu}(\partial_{\mu}e_{c\nu}-\Gamma_{\nu\mu}^{\alpha}e_{c\alpha})\end{aligned} (20)

using a frame field eaμe_{a}^{\mu}, which satisfies

eaμebνgμν=δab.\displaystyle e_{a}^{\mu}e_{b}^{\nu}g_{\mu\nu}=\delta_{ab}\,. (21)

The covariant derivative satisfies a relation,

(Γaa)2=214,\displaystyle(\Gamma^{a}\nabla_{a})^{2}=\nabla^{2}-\frac{1}{4}\mathcal{R}\,, (22)

where \mathcal{R} is a Ricci scalar.

For d\mathbb{H}^{d}, a solution of the equation of motion for the massive fermion behaves

ψzΔ±\displaystyle\psi\sim z^{\Delta_{\pm}} (23)

near the boundary at z=0z=0, where Δ±\Delta_{\pm} is given by

Δ±d12=±MR.\displaystyle\Delta_{\pm}-\frac{d-1}{2}=\pm MR\,. (24)

Then, Δ±\Delta_{\pm} become degenerate in the massless limit, and the allowed asymptotic behaviour is unique for a massless fermion.

For the product space p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}, we consider a massless fermion from the beginning. We first decompose the fermionic field as

ψ(z,x,θ)=ψp+1(z,x)ψ,𝕊q1(θ),\displaystyle\psi(z,x,\theta)=\sum_{\ell}\psi_{\mathbb{H}^{p+1}}(z,x)\otimes\psi_{\ell,\mathbb{S}^{q-1}}(\theta)\,, (25)

and the gamma matrices are also decomposed similarly. For even pp and even q4q\geq 4, the decomposition of the fermion (25) has an additional U(1)U(1) as is clear from that the rank of the spinor representation are different in both sides of (25).444We are indebted to D. Rodriguez-Gomez and J. G. Russo for the decomposition of the fermion. Mathematically, this is equivalent to a decomposition of a representation of SO(p+q2)SO(\frac{p+q}{2}) to that of SO(p2)×SO(q22)×U(1)SO(\frac{p}{2})\times SO(\frac{q-2}{2})\times U(1). The spherical part ψ,𝕊q1(θ)\psi_{\ell,\mathbb{S}^{q-1}}(\theta) satisfies the equation,

(Γ)𝕊q1ψ,𝕊q1(θ)=±i+q12Rψ,𝕊q1(θ),\displaystyle(\Gamma\cdot\nabla)_{\mathbb{S}^{q-1}}\psi_{\ell,\mathbb{S}^{q-1}}(\theta)=\pm\,\textrm{i}\,\frac{\ell+\frac{q-1}{2}}{R}\psi_{\ell,\mathbb{S}^{q-1}}(\theta)\,, (26)

where we decomposed the covariant derivative on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} into the covariant derivatives on p+1\mathbb{H}^{p+1} and 𝕊q1\mathbb{S}^{q-1} appropriately. Then, the equation of motion reduces to

(i(Γ)p+1±+q12R)ψp+1(z,x)=0.\displaystyle\left(\textrm{i}\,(\Gamma\cdot\nabla)_{\mathbb{H}^{p+1}}\pm\frac{\ell+\frac{q-1}{2}}{R}\right)\psi_{\mathbb{H}^{p+1}}(z,x)=0\,. (27)

The solution of this equation of motion behaves as

ψp+1(z,x)zΔ±\displaystyle\psi_{\mathbb{H}^{p+1}}(z,x)\sim z^{\Delta_{\pm}^{\ell}} (28)

near the boundary, z=0z=0, and Δ±\Delta_{\pm}^{\ell} is given by

Δ±=p2±(+q12).\displaystyle\Delta_{\pm}^{\ell}=\frac{p}{2}\pm\left(\ell+\frac{q-1}{2}\right)\,. (29)

The parameter Δ±\Delta_{\pm}^{\ell} can be understood as the conformal dimensions of operators localising on a pp-dimensional conformal defect at the boundary of p+1\mathbb{H}^{p+1}. Then not all the operators with conformal dimensions (29) are allowed to exist due to the unitarity bound in pp dimensions Minwalla:1997ka :

Δp12.\displaystyle\Delta\geq\frac{p-1}{2}\,. (30)

Δ+\Delta_{+}^{\ell} is always above the unitarity bound, while Δ\Delta_{-}^{\ell} is not necessary to satisfy the unitarity bound unless

1q2.\displaystyle\ell\leq 1-\frac{q}{2}\ . (31)

Hence the mode with =0\ell=0 for q=2q=2 is allowed to have the boundary conditions corresponding to Δ\Delta_{-}^{\ell}.

The allowed boundary conditions for the massless fermion are classified as follows and are listed in table 1.

q=1q=1 case:

In this case, the geometry is the hyperbolic space d\mathbb{H}^{d}, and the allowed asymptotic behaviour is unique,

Mixed b.c.:Δ=d12.\displaystyle\text{Mixed b.c.}:\quad\Delta=\frac{d-1}{2}\,. (32)

This boundary condition is called a mixed boundary condition.

q=2q=2 case:

Only the =0\ell=0 mode is allowed, resulting in a nontrivial boundary condition with Δ=0=p12\Delta^{\ell=0}_{-}=\frac{p-1}{2} for p2p\geq 2. Note that Δ=0\Delta^{\ell=0}_{-} saturates the unitarity bound. Then, two different boundary conditions are allowed,

Dirichlet b.c.:ΔD=Δ+for all ,Neumann b.c.:ΔN={Δfor =0,Δ+for 0.\displaystyle\begin{aligned} \text{Dirichlet b.c.}:&\quad\Delta_{\text{D}}=\Delta_{+}^{\ell}\quad\text{for all }\ell\,,\\ \text{Neumann b.c.}:&\quad\Delta_{\text{N}}=\begin{dcases}\Delta_{-}^{\ell}&\quad\text{for }\ell=0\,,\\ \Delta_{+}^{\ell}&\quad\text{for }\ell\neq 0\,.\end{dcases}\end{aligned} (33)

In particular, Δ=0\Delta^{\ell=0}_{-} vanishes for p=1p=1 case, and this implies that the defect operator is the identity operator. Thus, we exclude p=1p=1 case of a nontrivial boundary condition.

q3q\geq 3 case:

Only the Dirichlet type boundary condition is allowed,

Dirichlet b.c.:ΔD=Δ+for all .\displaystyle\text{Dirichlet b.c.}:\quad\Delta_{\text{D}}=\Delta_{+}^{\ell}\quad\text{for all }\ell\,. (34)
q=1q=1 q=2q=2 q=3q=3 q=4q=4 q=5q=5 \cdots
p=1p=1 Δ\Delta ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}} \cdots
p=2p=2 Δ\Delta ΔD/Δ=0\Delta_{\text{D}}/\Delta_{-}^{\ell=0} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}}
p=3p=3 Δ\Delta ΔD/Δ=0\Delta_{\text{D}}/\Delta_{-}^{\ell=0} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}} \cdots
p=4p=4 Δ\Delta ΔD/Δ=0\Delta_{\text{D}}/\Delta_{-}^{\ell=0} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}}
p=5p=5 Δ\Delta ΔD/Δ=0\Delta_{\text{D}}/\Delta_{-}^{\ell=0} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}} ΔD\Delta_{\text{D}}
\vdots \vdots \vdots \ddots
Table 1: Classification of the allowed boundary conditions in the free massless fermion. The Neumann boundary conditions exist in the shaded cells and the allowed modes differ from the Dirichlet ones are shown in the right side. For q=1q=1, the boundary condition is unique. For q=2q=2, Δ=0\Delta_{-}^{\ell=0} saturates the unitarity bound (30).

3 Free energy on 𝕊d\mathbb{S}^{d} and 𝕊d\mathbb{HS}^{d}

In this section, we compute free energies of a free massless Dirac fermion on 𝕊d\mathbb{S}^{d} and 𝕊d\mathbb{HS}^{d} using a zeta-function regularisation for a warm-up of the next section.

3.1 Free energy on 𝕊d\mathbb{S}^{d}

For a massless fermion on 𝕊d\mathbb{S}^{d}, the free energy is given by555As noted in Nishioka:2021uef ; Monin:2016bwf , there is an ambiguity to decompose the logarithmic function into two parts. We require two conditions to remove the ambiguity: the free energy in the Schwinger representation should be convergent and the resulting zeta function does not depend on the cutoff scale Λ~\tilde{\Lambda}. In (35), the two decomposed logarithmic functions are the same and the free energy in the Schwinger representation (40) is convergent in the \ell\to\infty limit.

F[𝕊d]=12trlog[Λ~2(Γaa)2]==0g(d)()log(ν(d)Λ~R),\displaystyle\begin{aligned} F[\mathbb{S}^{d}]&=-\frac{1}{2}\text{tr}\log\left[-\tilde{\Lambda}^{-2}(\Gamma^{a}\nabla_{a})^{2}\right]\\ &=-\sum_{\ell=0}^{\infty}g^{(d)}(\ell)\,\log\left(\frac{\nu_{\ell}^{(d)}}{\tilde{\Lambda}R}\right)\,,\end{aligned} (35)

where we use the equation,

Γaaψ=±iν(d)Rψ\displaystyle\Gamma^{a}\nabla_{a}\psi_{\ell}=\pm\,\textrm{i}\,\frac{\nu_{\ell}^{(d)}}{R}\psi_{\ell} (36)

with the eigenvalues,

ν(d)=+d2,=0,1,2,.\displaystyle\nu_{\ell}^{(d)}=\ell+\frac{d}{2}\,,\qquad\ell=0,1,2,\cdots\,. (37)

The degeneracy for each sign is given by

g±(d)()=rdΓ(+d)Γ(d)Γ(+1),\displaystyle g_{\pm}^{(d)}(\ell)=\frac{r_{d}\Gamma(\ell+d)}{\Gamma(d)\Gamma(\ell+1)}\,, (38)

where rdr_{d} (18) is the rank of the gamma matrix as before. For later convenience, we introduce a notation

g(d)()=2g±(d)()=2rdΓ(+d)Γ(d)Γ(+1).\displaystyle g^{(d)}(\ell)=2g_{\pm}^{(d)}(\ell)=\frac{2r_{d}\Gamma(\ell+d)}{\Gamma(d)\Gamma(\ell+1)}\,. (39)

We write the free energy (35) in the Schwinger representation,

F[𝕊d]=0dtt=0g(d)()etν(d)/(Λ~R),\displaystyle F[\mathbb{S}^{d}]=\int_{0}^{\infty}\,\frac{\mathrm{d}t}{t}\,\sum_{\ell=0}^{\infty}g^{(d)}(\ell)\,\mathrm{e}^{-t\nu_{\ell}^{(d)}/(\tilde{\Lambda}R)}\,, (40)

and we introduce the regularised free energy Vassilevich:2003xt to remove the divergence in the integral,

Fs[𝕊d]=0dtt1s=0g(d)()etν(d)/(Λ~R)=12(Λ~R)sΓ(s)ζ𝕊d(s),\displaystyle\begin{aligned} F_{s}[\mathbb{S}^{d}]&=\int_{0}^{\infty}\,\frac{\mathrm{d}t}{t^{1-s}}\,\sum_{\ell=0}^{\infty}g^{(d)}(\ell)\,\mathrm{e}^{-t\nu_{\ell}^{(d)}/(\tilde{\Lambda}R)}\\ &=\frac{1}{2}(\tilde{\Lambda}R)^{s}\,\Gamma(s)\,\zeta_{\mathbb{S}^{d}}(s)\,,\end{aligned} (41)

where the zeta function ζ𝕊d(s)\zeta_{\mathbb{S}^{d}}(s) is defined by

ζ𝕊d(s)2=0g(d)()(ν(d))s.\displaystyle\zeta_{\mathbb{S}^{d}}(s)\equiv 2\sum_{\ell=0}^{\infty}g^{(d)}(\ell)\,\left(\nu_{\ell}^{(d)}\right)^{-s}\ . (42)

Then the (unregularised) free energy is obtained in the s0s\to 0 limit:

Fs[𝕊d]=12(1sγE+log(Λ~R))ζ𝕊d(0)+12sζ𝕊d(0)+𝒪(s),\displaystyle F_{s}[\mathbb{S}^{d}]=\frac{1}{2}\left(\frac{1}{s}-\gamma_{\text{E}}+\log(\tilde{\Lambda}R)\right)\,\zeta_{\mathbb{S}^{d}}(0)+\frac{1}{2}\partial_{s}\zeta_{\mathbb{S}^{d}}(0)+\mathcal{O}(s)\ , (43)

which is divergent due to the pole at s=0s=0. Here γE\gamma_{\text{E}} is the Euler constant. By removing the pole, the remaining part becomes the renormalized free energy

Fren[𝕊d]12sζ𝕊d(0)+12log(ΛR)ζ𝕊d(0),\displaystyle F_{\text{ren}}[\mathbb{S}^{d}]\equiv\frac{1}{2}\partial_{s}\zeta_{\mathbb{S}^{d}}(0)+\frac{1}{2}\log(\Lambda R)\,\zeta_{\mathbb{S}^{d}}(0)\ , (44)

where Λ=eγEΛ~\Lambda=\mathrm{e}^{-\gamma_{\text{E}}}\,\tilde{\Lambda}.

To compute the zeta function, we expand the gamma functions in the degeneracy (39),

Γ(ν(d)+d2)Γ(ν(d)d2+1)={n=0d2(1)d2+nαn,d+1(ν(d))2n1d:even,n=0d12(1)d12+nβn,d+1(ν(d))2nd:odd,\displaystyle\frac{\Gamma\left(\nu_{\ell}^{(d)}+\frac{d}{2}\right)}{\Gamma\left(\nu_{\ell}^{(d)}-\frac{d}{2}+1\right)}=\begin{dcases}\sum_{n=0}^{\frac{d}{2}}(-1)^{\frac{d}{2}+n}\,\alpha_{n,d+1}\,\left(\nu_{\ell}^{(d)}\right)^{2n-1}&d:\text{even}\,,\\ \sum_{n=0}^{\frac{d-1}{2}}(-1)^{\frac{d-1}{2}+n}\,\beta_{n,d+1}\,\left(\nu_{\ell}^{(d)}\right)^{2n}&d:\text{odd}\,,\end{dcases} (45)

where we used the awkward suffix for αn,d+1\alpha_{n,d+1} and βn,d+1\beta_{n,d+1} to use the same notation in the scalar case Nishioka:2021uef . Since α0,d+1=0\alpha_{0,d+1}=0, we can omit the n=0n=0 term in the summation for even dd whenever we want. Using the asymptotic expansion ((5.11.14) in NIST:DLMF )

Γ(x+a)Γ(x+b)=k=0(x+a+b12)ab2k(ab2k)B2k(ab+1)(ab+12),\displaystyle\frac{\Gamma(x+a)}{\Gamma(x+b)}=\sum_{k=0}^{\infty}\,\left(x+\frac{a+b-1}{2}\right)^{a-b-2k}\,\binom{a-b}{2k}\,B_{2k}^{(a-b+1)}\left(\frac{a-b+1}{2}\right)\ , (46)

where Bk(m)(x)B_{k}^{(m)}(x) is the generalized Bernoulli polynomial, and comparing both sides, we find

Even d:αn,d+1\displaystyle\text{Even }d:\qquad\alpha_{n,d+1} =(1)d2+n(d1d2n)Bd2n(d)(d2),\displaystyle=(-1)^{\frac{d}{2}+n}\,\binom{d-1}{d-2n}\,B^{(d)}_{d-2n}\left(\frac{d}{2}\right)\,, (47)
Odd d:βn,d+1\displaystyle\text{Odd }d:\qquad\beta_{n,d+1} =(1)d12+n(d1d12n)Bd12n(d)(d2).\displaystyle=(-1)^{\frac{d-1}{2}+n}\,\binom{d-1}{d-1-2n}\,B^{(d)}_{d-1-2n}\left(\frac{d}{2}\right)\,. (48)

3.1.1 Odd dd

When dd is odd, the zeta function reduces to

ζ𝕊d(s)=4rdΓ(d)n=0d12(1)d12+nβn,d+1ζH(s2n,12),\displaystyle\zeta_{\mathbb{S}^{d}}(s)=\frac{4r_{d}}{\Gamma(d)}\sum_{n=0}^{\frac{d-1}{2}}(-1)^{\frac{d-1}{2}+n}\beta_{n,d+1}\,\zeta_{\text{H}}\left(s-2n,\frac{1}{2}\right)\,, (49)

where we use the identity

ζH(s2n,d2)=ζH(s2n,12)m=0d12(m+12)2ns\displaystyle\zeta_{\text{H}}\left(s-2n,\frac{d}{2}\right)=\zeta_{\text{H}}\left(s-2n,\frac{1}{2}\right)-\sum_{m=0}^{\frac{d-1}{2}}\left(m+\frac{1}{2}\right)^{2n-s} (50)

and the expansion (45) with the replacement ν(d)m+1/2\nu_{\ell}^{(d)}\to m+1/2. Since the zeta function at s=0s=0 vanishes,

ζ𝕊d(0)=4rdΓ(d)n=0d12(1)d12+nβn,d+1ζH(2n,12)=0,\displaystyle\begin{aligned} \zeta_{\mathbb{S}^{d}}(0)&=\frac{4r_{d}}{\Gamma(d)}\sum_{n=0}^{\frac{d-1}{2}}(-1)^{\frac{d-1}{2}+n}\beta_{n,d+1}\,\zeta_{\text{H}}\left(-2n,\frac{1}{2}\right)\\ &=0\,,\end{aligned} (51)

due to the fact ζH(2n,1/2)=0\zeta_{\text{H}}\left(-2n,1/2\right)=0, there is no conformal anomaly, and the finite part remains in the free energy:

Fren[𝕊d]=12sζ𝕊d(0)=rdΓ(d)(1)d+12β0,d+1log2+2rdΓ(d)n=1d12(1)d12+nβn,d+1(22n1)ζ(2n).\displaystyle\begin{aligned} F_{\text{ren}}[\mathbb{S}^{d}]&=\frac{1}{2}\partial_{s}\zeta_{\mathbb{S}^{d}}(0)\\ &=\frac{r_{d}}{\Gamma(d)}(-1)^{\frac{d+1}{2}}\beta_{0,d+1}\,\log 2+\frac{2r_{d}}{\Gamma(d)}\sum_{n=1}^{\frac{d-1}{2}}(-1)^{\frac{d-1}{2}+n}\beta_{n,d+1}\,(2^{-2n}-1)\zeta^{\prime}(-2n)\,.\end{aligned} (52)

This formula correctly reproduces the known results Marino:2011nm ; Klebanov:2011gs . The finite parts of the free energy for d9d\leq 9 are listed in table 3 in appendix A.

3.1.2 Even dd

When dd is even the zeta function is given by

ζ𝕊d(s)=4rdΓ(d)n=0d2(1)d2+nαn,d+1ζ(s2n+1),\displaystyle\zeta_{\mathbb{S}^{d}}(s)=\frac{4r_{d}}{\Gamma(d)}\sum_{n=0}^{\frac{d}{2}}(-1)^{\frac{d}{2}+n}\,\alpha_{n,d+1}\,\zeta(s-2n+1)\,, (53)

where we use the identity

ζH(s2n+1,d2)=ζ(s2n+1)m=0d22(m+1)2n1s\displaystyle\begin{aligned} \zeta_{\text{H}}\left(s-2n+1,\frac{d}{2}\right)=\zeta(s-2n+1)-\sum_{m=0}^{\frac{d}{2}-2}\left(m+1\right)^{2n-1-s}\end{aligned} (54)

and the expansion (45) with the replacement ν(d)m+1\nu_{\ell}^{(d)}\to m+1. The renormalized free energy is given by

Fren[𝕊d]=A[𝕊d]log(ΛR)+Ffin[𝕊d],\displaystyle F_{\text{ren}}[\mathbb{S}^{d}]=-A[\mathbb{S}^{d}]\,\log(\Lambda R)+F_{\text{fin}}[\mathbb{S}^{d}]\ , (55)

with the anomaly coefficient

A[𝕊d]=12ζ𝕊d(0)=rdΓ(d)n=1d2(1)d2+nαn,d+1B2nn,\displaystyle\begin{aligned} A[\mathbb{S}^{d}]&=-\frac{1}{2}\zeta_{\mathbb{S}^{d}}(0)\\ &=\frac{r_{d}}{\Gamma(d)}\sum_{n=1}^{\frac{d}{2}}(-1)^{\frac{d}{2}+n}\,\alpha_{n,d+1}\,\frac{B_{2n}}{n}\,,\end{aligned} (56)

where B2nB_{2n} is the Bernoulli number, and the finite part

Ffin[𝕊d]=12sζ𝕊d(0)=2rdΓ(d)n=1d2(1)d2+nαn,d+1ζ(2n+1).\displaystyle\begin{aligned} F_{\text{fin}}[\mathbb{S}^{d}]&=\frac{1}{2}\partial_{s}\zeta_{\mathbb{S}^{d}}(0)\\ &=\frac{2r_{d}}{\Gamma(d)}\sum_{n=1}^{\frac{d}{2}}(-1)^{\frac{d}{2}+n}\,\alpha_{n,d+1}\,\zeta^{\prime}(-2n+1)\,.\end{aligned} (57)

There is a logarithmic divergent term associated with the conformal anomaly in the free energy. The free energy (55) with (56) correctly reproduces the known conformal anomaly Giombi:2014xxa . The anomaly coefficients and the finite parts of the free energy for d10d\leq 10 are listed in tables 2 and 3 in appendix A.

3.1.3 Interpolating aa and FF

The finite part of the free energy (52) and the anomaly coefficient of the free energy (56) do not depend on the choice of the cutoff scale and are universal in this sense. Thus, we introduce the “universal” free energy:

Funiv[𝕊d]={Ffin[𝕊d]d:odd,A[𝕊d]log(Rϵ)d:even,\displaystyle F_{\text{univ}}[\mathbb{S}^{d}]=\begin{dcases}F_{\text{fin}}[\mathbb{S}^{d}]&\qquad d:\text{odd}\,,\\ -A[\mathbb{S}^{d}]\,\log\left(\frac{R}{\epsilon}\right)&\qquad d:\text{even}\,,\end{dcases} (58)

where ϵ\epsilon is used for the cutoff instead of Λ\Lambda. In Giombi:2014xxa , it is pointed out that the universal free energy has an integral representation:

Funiv[𝕊d]=2rdsin(πd2)Γ(d+1)012ducos(πu)Γ(d+12+u)Γ(d+12u).\displaystyle F_{\text{univ}}[\mathbb{S}^{d}]=-\frac{2r_{d}}{\sin\left(\frac{\pi d}{2}\right)\Gamma(d+1)}\int_{0}^{\frac{1}{2}}\!\mathrm{d}u\,\cos\left(\pi u\right)\Gamma\left(\frac{d+1}{2}+u\right)\Gamma\left(\frac{d+1}{2}-u\right)\,. (59)

For odd dd, the prefactor of the universal free energy (59) is finite. However, for even dd, the prefactor in (59) is divergent due to the sine function. This divergence may be replaced with the logarithmic divergence by introducing a small cutoff ϵ\epsilon,

1sin(πd2)={(1)d+12d:odd,(1)d22πlog(Rϵ)d:even.\displaystyle-\frac{1}{\sin\left(\frac{\pi d}{2}\right)}=\begin{dcases}(-1)^{\frac{d+1}{2}}&\qquad d:\text{odd}\,,\\ (-1)^{\frac{d}{2}}\,\frac{2}{\pi}\,\log\left(\frac{R}{\epsilon}\right)&\qquad d:\text{even}\,.\end{dcases} (60)

Using the replacement (60), the universal free energy in the integral representation (59) has the same behaviour of (58). A proof of the equivalence of the two expressions (58) and (59) is presented in Dowker:2017cqe .

3.2 Free energy on 𝕊d\mathbb{HS}^{d}

Next, let us consider the free energy on the hemisphere. At the boundary of 𝕊d\mathbb{HS}^{d}, a mixed boundary condition is imposed Dowker:1995sw ,

P+ψ=0.\displaystyle P_{+}\psi=0\,. (61)

Here P+P_{+} is a projection operator,

P+=12(1iΓΓaeaμnμ)\displaystyle P_{+}=\frac{1}{2}\left(1-\textrm{i}\,\Gamma^{\ast}\Gamma^{a}e_{a}^{\mu}n_{\mu}\right) (62)

with a chirality matrix Γ\Gamma^{\ast} and an incoming normal vector nμn_{\mu}. See appendix A in Dowker:1995sw for the detail of a construction of the chirality matrix Γ\Gamma^{\ast}. The mixed boundary condition preserves a conformal symmetry.

A degeneracy with the mixed boundary condition is given by (38) which is just a half of the degeneracy of 𝕊d\mathbb{S}^{d} (39). Then, the free energy on 𝕊d\mathbb{HS}^{d} is nothing but half of that on 𝕊d\mathbb{S}^{d},

Fren[𝕊d]=12Fren[𝕊d].\displaystyle F_{\text{ren}}[\mathbb{HS}^{d}]=\frac{1}{2}F_{\text{ren}}[\mathbb{S}^{d}]\,. (63)

This formula correctly reproduces the known results Rodriguez-Gomez:2017aca .666In Rodriguez-Gomez:2017aca , anomaly coefficients of a ball 𝔹d\mathbb{B}^{d} which is conformally equivalent to d\mathbb{H}^{d}, is obtained.

4 Free energy on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} with Dirichlet boundary condition

In this section, we first compute a free energy of a massive fermion on d\mathbb{H}^{d}, although we are interested in a massless (conformal) fermion. This is because the angular momentum along 𝕊q1\mathbb{S}^{q-1} can be regarded as a mass on p+1\mathbb{H}^{p+1} due to a Kaluza-Klein mechanism when we compute the free energy on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}. After that, we compute a free energy of a massless fermion on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}.

4.1 Free energy on d\mathbb{H}^{d}

In this section, we extend a computation of the zeta function for a massive fermion on d\mathbb{H}^{d} for d=3,4d=3,4 in Bytsenko:1994bc to general dimensions.

The free energy of the massive Dirac fermion with mass M=m/RM=m/R is given by

F[d](m)=12trlog[Λ~2((Γaa)2M2)]=120dωμ(d)(ω)[log(ω+imΛ~R)+log(ωimΛ~R)],\displaystyle\begin{aligned} F[\mathbb{H}^{d}](m)&=-\frac{1}{2}\text{tr}\log\left[-\tilde{\Lambda}^{-2}\left((\Gamma^{a}\nabla_{a})^{2}-M^{2}\right)\right]\\ &=-\frac{1}{2}\int_{0}^{\infty}\!\mathrm{d}\omega\,\mu^{(d)}(\omega)\left[\log\left(\frac{\omega+\textrm{i}\,m}{\tilde{\Lambda}R}\right)+\log\left(\frac{\omega-\textrm{i}\,m}{\tilde{\Lambda}R}\right)\right]\,,\end{aligned} (64)

where Λ~\tilde{\Lambda} is the UV cutoff scale introduced to make the integral dimensionless. The parameter ω\omega is an eigenvalue of the equation,

Γaaψω=iωRψω,ω0,\displaystyle\Gamma^{a}\nabla_{a}\psi_{\omega}=\textrm{i}\,\frac{\omega}{R}\psi_{\omega}\,,\qquad\omega\geq 0\,, (65)

and the Plancherel measure of a fermion on d\mathbb{H}^{d} of unit radius takes the form Camporesi:1995fb

μ(d)(ω)=cdrd|Γ(d2+iω)Γ(12+iω)|2=cdrd{j=12d22(ω2+j2)d:odd,ωcoth(πω)j=1d22(ω2+j2)d:even,\displaystyle\begin{aligned} \mu^{(d)}(\omega)&=c_{d}\,r_{d}\,\left|\frac{\Gamma\left(\frac{d}{2}+\textrm{i}\,\omega\right)}{\Gamma\left(\frac{1}{2}+\textrm{i}\,\omega\right)}\right|^{2}\\ &=c_{d}\,r_{d}\,\begin{dcases}\prod_{j=\frac{1}{2}}^{\frac{d-2}{2}}(\omega^{2}+j^{2})&d:\text{odd}\,,\\ \omega\,\coth(\pi\omega)\,\prod_{j=1}^{\frac{d-2}{2}}(\omega^{2}+j^{2})&d:\text{even}\,,\end{dcases}\end{aligned} (66)

with the coefficient

cd=Vol(d)2d1πd/2Γ(d/2)\displaystyle c_{d}=\frac{\text{Vol}(\mathbb{H}^{d})}{2^{d-1}\pi^{d/2}\Gamma(d/2)} (67)

and the rank of the gamma matrix rdr_{d} (18). The regularised volume of the hyperbolic space is given by

Vol(d)=πd12Γ(1d2)=πd+12sin(πd12)Γ(d+12).\displaystyle\text{Vol}(\mathbb{H}^{d})=\pi^{\frac{d-1}{2}}\Gamma\left(\frac{1-d}{2}\right)=-\frac{\pi^{\frac{d+1}{2}}}{\sin\left(\pi\frac{d-1}{2}\right)\Gamma\left(\frac{d+1}{2}\right)}\,. (68)

The hyperbolic volume is finite for even dd but divergent for odd dd due to the pole of the sine function. By introducing a small cutoff parameter ϵ\epsilon, the sine function can be replaced by the logarithmic divergence,

1sin(πd12)={(1)d122πlog(Rϵ)d:odd,(1)d2d:even.\displaystyle-\frac{1}{\sin\left(\pi\,\frac{d-1}{2}\right)}=\begin{dcases}(-1)^{\frac{d-1}{2}}\frac{2}{\pi}\,\log\left(\frac{R}{\epsilon}\right)&\qquad d:\text{odd}\,,\\ (-1)^{\frac{d}{2}}&\qquad d:\text{even}\,.\end{dcases} (69)

Then, the coefficient becomes

cd=1sin(πd12)Γ(d)=1Γ(d){(1)d122πlog(Rϵ)d:odd,(1)d2d:even.\displaystyle c_{d}=-\frac{1}{\sin\left(\pi\frac{d-1}{2}\right)\Gamma(d)}=\frac{1}{\Gamma(d)}\begin{dcases}(-1)^{\frac{d-1}{2}}\,\frac{2}{\pi}\,\log\left(\frac{R}{\epsilon}\right)&\qquad d:\text{odd}\,,\\ (-1)^{\frac{d}{2}}&\qquad d:\text{even}\,.\end{dcases} (70)

Since the Plancherel measure (66) needs to satisfy the square integrability condition, the free energy is defined for m0m\geq 0. That is, equation (64) represents a free energy with a boundary condition, Δ=(d1)/2+m\Delta=(d-1)/2+m. We add a mass term in the free energy (64) for the regularisation of the zero mode, and we take a massless limit to obtain free energies of the conformal fermion.

By using the Schwinger representation, the free energy can be written as

F[d](m)=120dtt0dωμ(d)(ω)(et(ω+im)/Λ~R+et(ωim)/Λ~R).\displaystyle F[\mathbb{H}^{d}](m)=\frac{1}{2}\int_{0}^{\infty}\!\frac{\mathrm{d}t}{t}\int_{0}^{\infty}\!\mathrm{d}\omega\,\mu^{(d)}(\omega)\,\left(\mathrm{e}^{-t(\omega+\textrm{i}\,m)/\tilde{\Lambda}R}+\mathrm{e}^{-t(\omega-\textrm{i}\,m)/\tilde{\Lambda}R}\right)\,. (71)

To remove the divergence in the integral, we introduce the regularised free energy

Fs[d](m)=120dtt1s0dωμ(d)(ω)(et(ω+im)/Λ~R+et(ωim)/Λ~R)=12(Λ~R)sΓ(s)ζd(s,m),\displaystyle\begin{aligned} F_{s}[\mathbb{H}^{d}](m)&=\frac{1}{2}\int_{0}^{\infty}\!\frac{\mathrm{d}t}{t^{1-s}}\int_{0}^{\infty}\!\mathrm{d}\omega\,\mu^{(d)}(\omega)\,\left(\mathrm{e}^{-t(\omega+\textrm{i}\,m)/\tilde{\Lambda}R}+\mathrm{e}^{-t(\omega-\textrm{i}\,m)/\tilde{\Lambda}R}\right)\\ &=\frac{1}{2}(\tilde{\Lambda}R)^{s}\Gamma(s)\zeta_{\mathbb{H}^{d}}(s,m)\,,\end{aligned} (72)

where the zeta function is defined as

ζd(s,m)=0dωμ(d)(ω)((ω+im)s+(ωim)s).\displaystyle\zeta_{\mathbb{H}^{d}}(s,m)=\int_{0}^{\infty}\!\mathrm{d}\omega\,\mu^{(d)}(\omega)\,\left((\omega+\textrm{i}\,m)^{-s}+(\omega-\textrm{i}\,m)^{-s}\right)\,. (73)

Then, the (unregularised) free energy is obtained in the s0s\to 0 limit,

Fs[d](m)=12(1sγE+log(Λ~R))ζd(s,m)+12sζd(s,m)+𝒪(s).\displaystyle F_{s}[\mathbb{H}^{d}](m)=\frac{1}{2}\left(\frac{1}{s}-\gamma_{\text{E}}+\log(\tilde{\Lambda}R)\right)\zeta_{\mathbb{H}^{d}}(s,m)+\frac{1}{2}\partial_{s}\zeta_{\mathbb{H}^{d}}(s,m)+\mathcal{O}(s)\,. (74)

By removing the pole at s=0s=0, we obtain the renormalized free energy

Fren[d](m)=12sζd(0,m)+12log(ΛR)ζd(0,m),\displaystyle F_{\text{ren}}[\mathbb{H}^{d}](m)=\frac{1}{2}\partial_{s}\zeta_{\mathbb{H}^{d}}(0,m)+\frac{1}{2}\log(\Lambda R)\zeta_{\mathbb{H}^{d}}(0,m)\,, (75)

where Λ=eγEΛ~\Lambda=\mathrm{e}^{-\gamma_{\text{E}}}\tilde{\Lambda}.

In the following, we will compute the renormalized free energy by evaluating the zeta function based on the method used in Camporesi:1994ga ; Bytsenko:1995ak .

4.1.1 Odd dd

Using the expansion of the Plancherel measure (66)

μ(d)(ω)=cdrdk=0d12βk,d+1ω2k,\displaystyle\mu^{(d)}(\omega)=c_{d}\,r_{d}\sum_{k=0}^{\frac{d-1}{2}}\beta_{k,d+1}\omega^{2k}\,, (76)

the zeta function is convergent for Res>2k+1\text{Re}\,s>2k+1,

ζd(s,m)=cdrdk=0d12βk,d+10dωω2k((ω+im)s+(ωim)s)=2cdrdk=0d12βk,d+1(1)ksin(πs2)m2k+1sΓ(2k+1)i=12k+11si.\displaystyle\begin{aligned} \zeta_{\mathbb{H}^{d}}(s,m)&=c_{d}\,r_{d}\sum_{k=0}^{\frac{d-1}{2}}\beta_{k,d+1}\int_{0}^{\infty}\!\mathrm{d}\omega\,\omega^{2k}\left((\omega+\textrm{i}\,m)^{-s}+(\omega-\textrm{i}\,m)^{-s}\right)\\ &=2c_{d}\,r_{d}\sum_{k=0}^{\frac{d-1}{2}}\beta_{k,d+1}(-1)^{k}\sin\left(\frac{\pi s}{2}\right)m^{2k+1-s}\Gamma(2k+1)\prod_{i=1}^{2k+1}\frac{1}{s-i}\,.\end{aligned} (77)

We immediately obtain

ζd(0,m)\displaystyle\zeta_{\mathbb{H}^{d}}(0,m) =0,\displaystyle=0\,, (78)
sζd(0,m)\displaystyle\partial_{s}\zeta_{\mathbb{H}^{d}}(0,m) =cdrdk=0d12βk,d+1πm2k+12k+1(1)k+1.\displaystyle=c_{d}\,r_{d}\sum_{k=0}^{\frac{d-1}{2}}\beta_{k,d+1}\frac{\pi m^{2k+1}}{2k+1}(-1)^{k+1}\,. (79)

For the massive fermion, the renormalized free energy is given by

Fren[d](m)=12sζd(0,m)=(2)d12Γ(d)k=0d12βk,d+1m2k+12k+1(1)k+1log(Rϵ).\displaystyle\begin{aligned} F_{\text{ren}}[\mathbb{H}^{d}](m)&=\frac{1}{2}\partial_{s}\zeta_{\mathbb{H}^{d}}(0,m)\\ &=\frac{(-2)^{\frac{d-1}{2}}}{\Gamma(d)}\sum_{k=0}^{\frac{d-1}{2}}\beta_{k,d+1}\frac{m^{2k+1}}{2k+1}(-1)^{k+1}\log\left(\frac{R}{\epsilon}\right)\,.\end{aligned} (80)

In the massless limit, the renormalized free energy vanishes,

Fren[d]=0.\displaystyle F_{\text{ren}}[\mathbb{H}^{d}]=0\,. (81)

Here and hereafter, we omit (0)(0) in free energies for the massless fermion. Equation (81) implies that the boundary anomaly does not exist.

4.1.2 Even dd

Using the expansion of the Plancherel (66)

μ(d)(ω)=cdrdcoth(πω)k=1d2αk,d+1ω2k1,\displaystyle\mu^{(d)}(\omega)=c_{d}\,r_{d}\coth(\pi\omega)\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}\omega^{2k-1}\,, (82)

and the identity

coth(πω)=1+2e2πω1,\displaystyle\coth(\pi\omega)=1+\frac{2}{\mathrm{e}^{2\pi\omega}-1}\,, (83)

the zeta function can be decomposed into two parts,

ζd(s,m)\displaystyle\zeta_{\mathbb{H}^{d}}(s,m) =ζd(1)(s,m)+ζd(2)(s,m),\displaystyle=\zeta_{\mathbb{H}^{d}}^{(1)}(s,m)+\zeta_{\mathbb{H}^{d}}^{(2)}(s,m)\,, (84)
ζd(1)(s,m)\displaystyle\zeta_{\mathbb{H}^{d}}^{(1)}(s,m) =cdrdk=1d2αk,d+10dωω2k1((ω+im)s+(ωim)s),\displaystyle=c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}\int_{0}^{\infty}\!\mathrm{d}\omega\,\omega^{2k-1}\left((\omega+\textrm{i}\,m)^{-s}+(\omega-\textrm{i}\,m)^{-s}\right)\,, (85)
ζd(2)(s,m)\displaystyle\zeta_{\mathbb{H}^{d}}^{(2)}(s,m) =2cdrdk=1d2αk,d+10dωω2k1e2πω1((ω+im)s+(ωim)s).\displaystyle=2c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\,\left((\omega+\textrm{i}\,m)^{-s}+(\omega-\textrm{i}\,m)^{-s}\right)\,. (86)

The first term of the zeta function can be calculated as

ζd(1)(s,m)=2cdrdk=1d2αk,d+1m2ks(1)kcos(πs2)Γ(2k)i=12k1si,\displaystyle\zeta_{\mathbb{H}^{d}}^{(1)}(s,m)=2c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}m^{2k-s}(-1)^{k}\cos\left(\frac{\pi s}{2}\right)\Gamma(2k)\prod_{i=1}^{2k}\frac{1}{s-i}\,, (87)

and we can easily read off

ζd(1)(0,m)\displaystyle\zeta_{\mathbb{H}^{d}}^{(1)}(0,m) =cdrdk=1d2αk,d+1(1)kkm2k,\displaystyle=c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}\frac{(-1)^{k}}{k}m^{2k}\,, (88)
sζd(1)(0,m)\displaystyle\partial_{s}\zeta_{\mathbb{H}^{d}}^{(1)}(0,m) =cdrdk=1d2αk,d+1(1)kkm2k(H2klogm),\displaystyle=c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}\frac{(-1)^{k}}{k}m^{2k}\left(H_{2k}-\log m\right)\,, (89)

where H2kH_{2k} is a harmonic number.

On the other hand, it is difficult to perform the integral in ζd(2)(s,m)\zeta_{\mathbb{H}^{d}}^{(2)}(s,m), hence we compute ζd(2)(0,m)\zeta_{\mathbb{H}^{d}}^{(2)}(0,m) and sζd(2)(0,m)\partial_{s}\zeta_{\mathbb{H}^{d}}^{(2)}(0,m) instead. ζd(2)(0,m)\zeta_{\mathbb{H}^{d}}^{(2)}(0,m) is independent on the mass term,

ζd(2)(0,m)=cdrdk=1d2αk,d+1(1)k+1B2kk,\displaystyle\zeta_{\mathbb{H}^{d}}^{(2)}(0,m)=c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}(-1)^{k+1}\frac{B_{2k}}{k}\,, (90)

and the derivative can be computed as

sζd(2)(0,m)=2cdrdk=1d2αk,d+1fk(m)\displaystyle\partial_{s}\zeta_{\mathbb{H}^{d}}^{(2)}(0,m)=-2c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}f_{k}(m) (91)

with

fk(m)=0dωω2k1e2πω1log(ω2+m2)=(1)k[m2k12(2k1)+m2k4k(1klog(m2))+l=1k1B2l4lm2k2lklB2kH2k12k+r=02k1(1)r(2k1r)m2k1r(ζ(r,m)Br+1(m)Hrr+1)],\displaystyle\begin{aligned} f_{k}(m)&=\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\log(\omega^{2}+m^{2})\\ &=(-1)^{k}\left[\frac{m^{2k-1}}{2(2k-1)}+\frac{m^{2k}}{4k}\left(\frac{1}{k}-\log(m^{2})\right)+\sum_{l=1}^{k-1}\frac{B_{2l}}{4l}\frac{m^{2k-2l}}{k-l}-\frac{B_{2k}H_{2k-1}}{2k}\right.\\ &\left.\phantom{(-1)^{k}}\qquad+\sum_{r=0}^{2k-1}(-1)^{r}\binom{2k-1}{r}m^{2k-1-r}\left(\zeta^{\prime}(-r,m)-\frac{B_{r+1}(m)H_{r}}{r+1}\right)\right]\,,\end{aligned} (92)

where Br+1(m)B_{r+1}(m) is a Bernoulli polynomial. See appendix B for the detailed derivation of (92).

We obtain the renormalized free energy for the massive fermion,

Fren[d](m)=A[d](m)log(ΛR)+Ffin[d](m)\displaystyle F_{\text{ren}}[\mathbb{H}^{d}](m)=-A[\mathbb{H}^{d}](m)\log(\Lambda R)+F_{\text{fin}}[\mathbb{H}^{d}](m) (93)

with the coefficient of the logarithmic divergent part

A[d](m)=12(ζd(1)(0,m)+ζd(2)(0,m))=cdrdk=1d2αk,d+1(1)kB2km2k2k,\displaystyle\begin{aligned} A[\mathbb{H}^{d}](m)&=-\frac{1}{2}\left(\zeta_{\mathbb{H}^{d}}^{(1)}(0,m)+\zeta_{\mathbb{H}^{d}}^{(2)}(0,m)\right)\\ &=c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}(-1)^{k}\frac{B_{2k}-m^{2k}}{2k}\,,\end{aligned} (94)

and the finite part

Ffin[d](m)=12(sζd(1)(0,m)+sζd(2)(0,m))=cdrdk=1d2αk,d+1(1)k+1(m2kH2k12kB2kH2k12k+m2k12(2k1)+l=1k1B2l4lm2k2lkl+r=02k1(1)r(2k1r)m2k1r(ζ(r,m)Br+1(m)Hrr+1)).\displaystyle\begin{aligned} &F_{\text{fin}}[\mathbb{H}^{d}](m)=\frac{1}{2}\left(\partial_{s}\zeta_{\mathbb{H}^{d}}^{(1)}(0,m)+\partial_{s}\zeta_{\mathbb{H}^{d}}^{(2)}(0,m)\right)\\ &\qquad=c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}(-1)^{k+1}\left(-m^{2k}\frac{H_{2k-1}}{2k}-\frac{B_{2k}H_{2k-1}}{2k}+\frac{m^{2k-1}}{2(2k-1)}\right.\\ &\left.\qquad\quad+\sum_{l=1}^{k-1}\frac{B_{2l}}{4l}\frac{m^{2k-2l}}{k-l}+\sum_{r=0}^{2k-1}(-1)^{r}\binom{2k-1}{r}m^{2k-1-r}\left(\zeta^{\prime}(-r,m)-\frac{B_{r+1}(m)H_{r}}{r+1}\right)\right)\,.\end{aligned} (95)

In the massless limit, m0m\to 0, the renormalized free energy has a simple expression

Fren[d]=A[d]log(ΛR)+Ffin[d],\displaystyle F_{\text{ren}}[\mathbb{H}^{d}]=-A[\mathbb{H}^{d}]\log(\Lambda R)+F_{\text{fin}}[\mathbb{H}^{d}]\,, (96)

with the anomaly coefficient and the finite part

A[d]\displaystyle A[\mathbb{H}^{d}] =cdrdk=1d2αk,d+1(1)k+1B2k2k,\displaystyle=c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}(-1)^{k+1}\frac{B_{2k}}{2k}\,, (97)
Ffin[d]\displaystyle F_{\text{fin}}[\mathbb{H}^{d}] =cdrdk=1d2αk,d+1(1)kζ(12k).\displaystyle=c_{d}\,r_{d}\sum_{k=1}^{\frac{d}{2}}\alpha_{k,d+1}(-1)^{k}\zeta^{\prime}(1-2k)\,. (98)

To see the massless limit of the finite term, it is convenient to use (152) instead of (92). The free energy on d\mathbb{H}^{d} is just the half of that on 𝕊d\mathbb{S}^{d},

Fren[d]=12Fren[𝕊d],\displaystyle F_{\text{ren}}[\mathbb{H}^{d}]=\frac{1}{2}F_{\text{ren}}[\mathbb{S}^{d}]\,, (99)

and this reproduces the known anomaly coefficients in literature Rodriguez-Gomez:2017aca .

The anomaly coefficients and the finite parts of the free energy for d10d\leq 10 are listed in tables 2 and 3 in appendix A.

4.2 Free energy on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}

The free energy on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} for the massless Dirac fermion except for even pp and even q4q\geq 4 is expressed as

F[p+1×𝕊q1]=12=0g(q1)()0dωμ(p+1)(ω)log(ω2+(ν(q1))2Λ~2R2),\displaystyle F[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]=-\frac{1}{2}\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\int_{0}^{\infty}\!\mathrm{d}\omega\,\mu^{(p+1)}(\omega)\log\left(\frac{\omega^{2}+\left(\nu_{\ell}^{(q-1)}\right)^{2}}{\tilde{\Lambda}^{2}R^{2}}\right)\,, (100)

with the Plancherel measure (66) and the degeneracy (39). For even pp and even q4q\geq 4, there are two fermions of opposite U(1)U(1) charge as the additional U(1)U(1) appeared in (25). Then, the free energy on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} with even pp and even q4q\geq 4 becomes

F[p+1×𝕊q1]==0g(q1)()0dωμ(p+1)(ω)log(ω2+(ν(q1))2Λ~2R2).\displaystyle F[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]=-\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\int_{0}^{\infty}\!\mathrm{d}\omega\,\mu^{(p+1)}(\omega)\log\left(\frac{\omega^{2}+\left(\nu_{\ell}^{(q-1)}\right)^{2}}{\tilde{\Lambda}^{2}R^{2}}\right)\,. (101)

To treat these two cases simultaneously, we introduce a notation,

dU(1)={2for even p and even q4,1otherwise.\displaystyle d_{U(1)}=\begin{dcases}2&\text{for even }p\text{ and even }q\geq 4\,,\\ 1&\text{otherwise}\,.\end{dcases} (102)

In the following, we compute the free energy using the zeta-function regularisation divided into two cases: even pp case and odd pp case. An important notice is that we use different decompositions for the logarithmic function depending on the evenness of pp.

4.2.1 Even pp

Performing similar computations in section 4.1.1, the renormalized free energy is given by

Fren[p+1×𝕊q1]=12ζp+1×𝕊q1(0)log(ΛR)+12sζp+1×𝕊q1(0)\displaystyle F_{\text{ren}}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]=\frac{1}{2}\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(0)\log(\Lambda R)+\frac{1}{2}\partial_{s}\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(0) (103)

with the zeta function

ζp+1×𝕊q1(s)=dU(1)=0g(q1)()ζp+1(s,ν(q1)).\displaystyle\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(s)=d_{U(1)}\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\zeta_{\mathbb{H}^{p+1}}\left(s,\nu_{\ell}^{(q-1)}\right)\,. (104)

Here ζp+1(s,ν(q1))\zeta_{\mathbb{H}^{p+1}}\left(s,\nu_{\ell}^{(q-1)}\right) is the zeta function of d\mathbb{H}^{d} with mass ν(q1)\nu_{\ell}^{(q-1)} (77). By using the expansion of the degeneracy (45) with the replacement dq1d\to q-1, the zeta function becomes

ζp+1×𝕊q1(s)=2dU(1)cp+1rp+1k=0p2βk,p+2(1)ksin(πs2)Γ(2k+1)i=12k+11si2rq1Γ(q1){n=0q12(1)q12+nαn,qζ(s2n2k)q:odd,n=0q22(1)q21+nβn,qζH(s2n2k1,12)q:even.\displaystyle\begin{aligned} \zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(s)&=2d_{U(1)}c_{p+1}\,r_{p+1}\sum_{k=0}^{\frac{p}{2}}\beta_{k,p+2}(-1)^{k}\sin\left(\frac{\pi s}{2}\right)\Gamma(2k+1)\prod_{i=1}^{2k+1}\frac{1}{s-i}\\ &\quad\cdot\frac{2r_{q-1}}{\Gamma(q-1)}\begin{dcases}\sum_{n=0}^{\frac{q-1}{2}}(-1)^{\frac{q-1}{2}+n}\,\alpha_{n,q}\,\zeta(s-2n-2k)&q:\text{odd}\,,\\ \sum_{n=0}^{\frac{q-2}{2}}(-1)^{\frac{q}{2}-1+n}\,\beta_{n,q}\,\zeta_{\text{H}}\left(s-2n-2k-1,\frac{1}{2}\right)&q:\text{even}\,.\end{dcases}\end{aligned} (105)

We immediately find

ζp+1×𝕊q1(0)=0,\displaystyle\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(0)=0\,, (106)

and

sζp+1×𝕊q1(0)=dU(1)cp+1rp+1k=0p2βk,p+2(1)kπ2k+12rq1Γ(q1){0q:odd,n=0q22(1)q21+nβn,qζH(2n2k1,12)q:even.\displaystyle\begin{aligned} \partial_{s}\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(0)&=-d_{U(1)}c_{p+1}\,r_{p+1}\sum_{k=0}^{\frac{p}{2}}\beta_{k,p+2}\frac{(-1)^{k}\pi}{2k+1}\\ &\cdot\frac{2r_{q-1}}{\Gamma(q-1)}\begin{dcases}0&q:\text{odd}\,,\\ \sum_{n=0}^{\frac{q-2}{2}}(-1)^{\frac{q}{2}-1+n}\,\beta_{n,q}\,\zeta_{\text{H}}\left(-2n-2k-1,\frac{1}{2}\right)&q:\text{even}\,.\end{dcases}\end{aligned} (107)

For even pp, we find the following:

  • For odd qq, the renormalized free energy vanishes,

    Fren[p+1×𝕊q1]=0.\displaystyle F_{\text{ren}}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]=0\,. (108)

    This implies that both bulk and defect anomalies vanish.

  • For even qq, the renormalized free energy has a defect anomaly which comes from the volume of p+1\mathbb{H}^{p+1},

    Fren[p+1×𝕊q1]=𝒜[p+1×𝕊q1]log(Rϵ)\displaystyle F_{\text{ren}}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]=-\mathcal{A}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]\log\left(\frac{R}{\epsilon}\right) (109)

    with

    𝒜[p+1×𝕊q1]=(2)p+q2dU(1)Γ(p+1)Γ(q1)k=0p2βk,p+2(1)k2k+1n=0q21(1)nβn,qζH(2n2k1,12).\displaystyle\begin{aligned} \mathcal{A}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]&=-\frac{(-2)^{\frac{p+q}{2}}d_{U(1)}}{\Gamma(p+1)\Gamma(q-1)}\,\sum_{k=0}^{\frac{p}{2}}\beta_{k,p+2}\frac{(-1)^{k}}{2k+1}\\ &\qquad\cdot\sum_{n=0}^{\frac{q}{2}-1}(-1)^{n}\,\beta_{n,q}\,\zeta_{\text{H}}\left(-2n-2k-1,\frac{1}{2}\right)\,.\end{aligned} (110)
  • For q=p+2q=p+2, the defect anomaly is proportional to the bulk anomaly on 𝕊2p+2\mathbb{S}^{2p+2},

    A[𝕊2p+2]=2𝒜[p+1×𝕊p+1],\displaystyle A[\mathbb{S}^{2p+2}]=2\mathcal{A}[\mathbb{H}^{p+1}\times\mathbb{S}^{p+1}]\,, (111)

    as in the scalar case Rodriguez-Gomez:2017kxf ; Nishioka:2021uef and the holographic case Rodriguez-Gomez:2017kxf . It is expected that the holographic result should be the same as the free fermion result because the anomaly coefficient does not depend on the strength of a coupling constant and the holography can be applied if the number of fermions is large.

For q=2q=2, our result correctly reproduces the free energy obtained in Klebanov:2011uf ; Beccaria:2017dmw (with an appropriate dimension of the spinor).

The anomaly coefficients of the free energy are listed in table 2 in appendix A.

4.2.2 Odd pp

In this section, we do not omit all equations in order to derive our main results of this section, (126) and (129). If the reader is not interested in the detail of the derivations, the reader can skip until (126).

For odd pp, the Plancherel measure (82) is decomposed into two parts using the identity (83), and the free energy consists of two parts,

F[p+1×𝕊q1]=cp+1rp+12k=1p+12αk,p+2=0g(q1)()0dωω2k1[log(ω+iν(q1)Λ~R)+log(ωiν(q1)Λ~R)]cp+1rp+1k=1p+12αk,p+20dωω2k1e2πω1=0g(q1)()[log(ν(q1)+iωΛ~R)+log(ν(q1)iωΛ~R)].\displaystyle\begin{aligned} F[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]&=-\frac{c_{p+1}\,r_{p+1}}{2}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\int_{0}^{\infty}\!\mathrm{d}\omega\,\omega^{2k-1}\\ &\qquad\qquad\cdot\left[\log\left(\frac{\omega+\textrm{i}\,\nu_{\ell}^{(q-1)}}{\tilde{\Lambda}R}\right)+\log\left(\frac{\omega-\textrm{i}\,\nu_{\ell}^{(q-1)}}{\tilde{\Lambda}R}\right)\right]\\ &\quad\phantom{=}-c_{p+1}\,r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\\ &\qquad\qquad\cdot\left[\log\left(\frac{\nu_{\ell}^{(q-1)}+\textrm{i}\,\omega}{\tilde{\Lambda}R}\right)+\log\left(\frac{\nu_{\ell}^{(q-1)}-\textrm{i}\,\omega}{\tilde{\Lambda}R}\right)\right]\,.\end{aligned} (112)

In contrast to section 4.1.2, we decompose the logarithmic function differently for each term, and the ordering of the summation over \ell and the integral over ω\omega is important.

The Schwinger representation of the free energy (112) is given by

F[p+1×𝕊q1]=cp+1rp+12k=1p+12αk,p+20dtt=0g(q1)()0dωω2k1[et(ω+iν(q1))/(Λ~R)+et(ωiν(q1))/(Λ~R)]+cp+1rp+1k=1p+12αk,p+20dtt0dωω2k1e2πω1=0g(q1)()[et(ν(q1)+iω)/(Λ~R)+et(ν(q1)iω)/(Λ~R)].\displaystyle\begin{aligned} F[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]&=\frac{c_{p+1}\,r_{p+1}}{2}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\int_{0}^{\infty}\!\frac{\mathrm{d}t}{t}\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\int_{0}^{\infty}\!\mathrm{d}\omega\,\omega^{2k-1}\\ &\qquad\qquad\cdot\left[\mathrm{e}^{-t(\omega+\textrm{i}\,\nu_{\ell}^{(q-1)})/(\tilde{\Lambda}R)}+\mathrm{e}^{-t(\omega-\textrm{i}\,\nu_{\ell}^{(q-1)})/(\tilde{\Lambda}R)}\right]\\ &\quad\phantom{=}+c_{p+1}\,r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\int_{0}^{\infty}\!\frac{\mathrm{d}t}{t}\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\\ &\qquad\qquad\cdot\left[\mathrm{e}^{-t(\nu_{\ell}^{(q-1)}+\textrm{i}\,\omega)/(\tilde{\Lambda}R)}+\mathrm{e}^{-t(\nu_{\ell}^{(q-1)}-\textrm{i}\,\omega)/(\tilde{\Lambda}R)}\right]\,.\end{aligned} (113)

For the first term, we first perform the integral over ω\omega, and hence the first term is convergent in the ω\omega\to\infty limit in the Schwinger representation. On the other hand, for the second term, we first perform the summation over \ell, and hence the second term is convergent in the \ell\to\infty limit in the Schwinger representation.

The renormalized free energy is given by

Fren[p+1×𝕊q1]=12ζp+1×𝕊q1(0)log(ΛR)+12sζp+1×𝕊q1(0),\displaystyle F_{\text{ren}}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]=\frac{1}{2}\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(0)\log(\Lambda R)+\frac{1}{2}\partial_{s}\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(0)\,, (114)

where the zeta function is a sum of two parts,

ζp+1×𝕊q1(s)=ζp+1×𝕊q1(1)(s)+ζp+1×𝕊q1(2)(s),\displaystyle\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}(s)=\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(1)}(s)+\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(2)}(s)\,, (115)

with

ζp+1×𝕊q1(1)(s)=cp+1rp+1k=1p+12αk,p+2=0g(q1)()0dωω2k1[(ω+iν(q1))s+(ωiν(q1))s],\displaystyle\begin{aligned} \zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(1)}(s)&=c_{p+1}\,r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\\ &\qquad\cdot\int_{0}^{\infty}\!\mathrm{d}\omega\,\omega^{2k-1}\left[\left(\omega+\textrm{i}\,\nu_{\ell}^{(q-1)}\right)^{-s}+\left(\omega-\textrm{i}\,\nu_{\ell}^{(q-1)}\right)^{-s}\right]\,,\end{aligned} (116)
ζp+1×𝕊q1(2)(s)=2cp+1rp+1k=1p+12αk,p+20dωω2k1e2πω1=0g(q1)()[(ν(q1)+iω)s+(ν(q1)iω)s].\displaystyle\begin{aligned} \zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(2)}(s)&=2c_{p+1}\,r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\\ &\qquad\cdot\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)\left[\left(\nu_{\ell}^{(q-1)}+\textrm{i}\,\omega\right)^{-s}+\left(\nu_{\ell}^{(q-1)}-\textrm{i}\,\omega\right)^{-s}\right]\,.\end{aligned} (117)

The first term in the zeta function is convergent for Res>2k\text{Re}\,s>2k,

ζp+1×𝕊q1(1)(s)=2cp+1rp+1k=1p+12αk,p+2(1)kcos(πs2)Γ(2k)i=12k1si2rq1Γ(q1){n=0q12(1)q12+nαn,qζ(s2n2k+1)q:odd,n=0q21(1)q21+nβn,qζH(s2n2k,12)q:even.\displaystyle\begin{aligned} \zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(1)}(s)&=2c_{p+1}r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}(-1)^{k}\cos\left(\frac{\pi s}{2}\right)\Gamma(2k)\prod_{i=1}^{2k}\frac{1}{s-i}\\ &\quad\cdot\frac{2r_{q-1}}{\Gamma(q-1)}\begin{dcases}\sum_{n=0}^{\frac{q-1}{2}}(-1)^{\frac{q-1}{2}+n}\,\alpha_{n,q}\,\zeta(s-2n-2k+1)&q:\text{odd}\,,\\ \sum_{n=0}^{\frac{q}{2}-1}(-1)^{\frac{q}{2}-1+n}\,\beta_{n,q}\,\zeta_{\text{H}}\left(s-2n-2k,\frac{1}{2}\right)&q:\text{even}\,.\end{dcases}\end{aligned} (118)

It follows that

ζp+1×𝕊q1(1)(0)=cp+1rp+12rq1Γ(q1)k=1p+12αk,p+2(1)kk{n=0q12(1)q12+nαn,qζ(2n2k+1)q:odd,0q:even,\displaystyle\begin{aligned} \zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(1)}(0)&=c_{p+1}\,r_{p+1}\frac{2r_{q-1}}{\Gamma(q-1)}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\frac{(-1)^{k}}{k}\\ &\quad\cdot\begin{dcases}\sum_{n=0}^{\frac{q-1}{2}}(-1)^{\frac{q-1}{2}+n}\,\alpha_{n,q}\,\zeta(-2n-2k+1)&q:\text{odd}\,,\\ 0&q:\text{even}\,,\end{dcases}\end{aligned} (119)

and

sζp+1×𝕊q1(1)(0)=cp+1rp+12rq1Γ(q1)k=1p+12αk,p+2(1)kk{n=0q12(1)q12+nαn,q[H2kζ(2n2k+1)+ζ(2n2k+1)]q:odd,n=0q21(1)q21+nβn,qsζH(2n2k,12)q:even.\displaystyle\begin{aligned} &\partial_{s}\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(1)}(0)\\ &\quad=c_{p+1}\,r_{p+1}\frac{2r_{q-1}}{\Gamma(q-1)}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\frac{(-1)^{k}}{k}\\ &\qquad\cdot\begin{dcases}\sum_{n=0}^{\frac{q-1}{2}}(-1)^{\frac{q-1}{2}+n}\,\alpha_{n,q}\,\left[H_{2k}\zeta(-2n-2k+1)+\zeta^{\prime}(-2n-2k+1)\right]&q:\text{odd}\,,\\ \sum_{n=0}^{\frac{q}{2}-1}(-1)^{\frac{q}{2}-1+n}\,\beta_{n,q}\,\partial_{s}\zeta_{\text{H}}\left(-2n-2k,\frac{1}{2}\right)&q:\text{even}\,.\end{dcases}\end{aligned} (120)

Using the expansion of the gamma functions in the degeneracy,

Γ(ν(q1)+q12)Γ(ν(q1)q12+1)={n=0q12(1)q12+nαn,qi=02n1(2n1i)(ν(q1)+iω)i(iω)2n1iq:odd,n=0q21(1)q21+nβn,qi=02n(2ni)(ν(q1)+iω)i(iω)2niq:even,\displaystyle\begin{aligned} &\frac{\Gamma\left(\nu_{\ell}^{(q-1)}+\frac{q-1}{2}\right)}{\Gamma\left(\nu_{\ell}^{(q-1)}-\frac{q-1}{2}+1\right)}\\ &\qquad=\begin{dcases}\sum_{n=0}^{\frac{q-1}{2}}(-1)^{\frac{q-1}{2}+n}\,\alpha_{n,q}\,\sum_{i=0}^{2n-1}\binom{2n-1}{i}\left(\nu_{\ell}^{(q-1)}+\textrm{i}\,\omega\right)^{i}(-\textrm{i}\,\omega)^{2n-1-i}&q:\text{odd}\,,\\ \sum_{n=0}^{\frac{q}{2}-1}(-1)^{\frac{q}{2}-1+n}\,\beta_{n,q}\,\sum_{i=0}^{2n}\binom{2n}{i}\left(\nu_{\ell}^{(q-1)}+\textrm{i}\,\omega\right)^{i}(-\textrm{i}\,\omega)^{2n-i}&q:\text{even}\,,\end{dcases}\end{aligned} (121)

and its complex conjugate, we obtain

ζp+1×𝕊q1(2)(s)=2cp+1rp+1k=1p+12αk,p+20dωω2k1e2πω12rq1Γ(q1){n=0q12(1)q12+nαn,qi=02n1(2n1i)(i)2n1iω2n1i[ζH(si,iω)+(1)i+1ζH(si,iω)]q:odd,n=0q21(1)q21+nβn,qi=02n(2ni)(i)2niω2ni[ζH(si,12+iω)+(1)iζH(si,12iω)]q:even.\displaystyle\begin{aligned} \zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(2)}(s)&=2c_{p+1}\,r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\frac{2r_{q-1}}{\Gamma(q-1)}\\ &\qquad\cdot\begin{dcases}\begin{aligned} &\sum_{n=0}^{\frac{q-1}{2}}(-1)^{\frac{q-1}{2}+n}\,\alpha_{n,q}\,\sum_{i=0}^{2n-1}\binom{2n-1}{i}(-\textrm{i})^{2n-1-i}\omega^{2n-1-i}\\ &\quad\cdot\left[\zeta_{\text{H}}\left(s-i,\textrm{i}\,\omega\right)+(-1)^{i+1}\zeta_{\text{H}}\left(s-i,-\textrm{i}\,\omega\right)\right]\end{aligned}&q:\text{odd}\,,\\ \begin{aligned} &\sum_{n=0}^{\frac{q}{2}-1}(-1)^{\frac{q}{2}-1+n}\,\beta_{n,q}\,\sum_{i=0}^{2n}\binom{2n}{i}(-\textrm{i})^{2n-i}\omega^{2n-i}\\ &\quad\cdot\left[\zeta_{\text{H}}\left(s-i,\frac{1}{2}+\textrm{i}\,\omega\right)+(-1)^{i}\zeta_{\text{H}}\left(s-i,\frac{1}{2}-\textrm{i}\,\omega\right)\right]\end{aligned}&q:\text{even}\,.\end{dcases}\end{aligned} (122)

Although it is difficult to perform the integral over ω\omega analytically, it is possible to simplify ζp+1×𝕊q1(2)(0)\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(2)}(0) and its derivative at s=0s=0 furthermore. After a bit of calculation, we obtain

ζp+1×𝕊q1(2)(0)={cp+1rp+12rq1Γ(q1)k=1p+12αk,p+2n=0q12(1)q12αn,qi=02n1(2n1i)(1)i+k+n+1i+1l=0i+12(i+12l)B2lB2k+2n2lk+nlq:odd,0q:even.\displaystyle\begin{aligned} \zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(2)}(0)=\begin{dcases}\begin{aligned} &c_{p+1}\,r_{p+1}\frac{2r_{q-1}}{\Gamma(q-1)}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\sum_{n=0}^{\frac{q-1}{2}}(-1)^{\frac{q-1}{2}}\,\alpha_{n,q}\\ &\quad\cdot\sum_{i=0}^{2n-1}\binom{2n-1}{i}\frac{(-1)^{i+k+n+1}}{i+1}\sum_{l=0}^{\lfloor\frac{i+1}{2}\rfloor}\binom{i+1}{2l}B_{2l}\frac{B_{2k+2n-2l}}{k+n-l}\end{aligned}&q:\text{odd}\,,\\ 0&q:\text{even}\,.\end{dcases}\end{aligned} (123)

In addition, the derivative of the zeta function at s=0s=0 can be computed as

sζp+1×𝕊q1(2)(0)=2cp+1rp+1k=1p+12αk,p+20dωω2k1e2πω12rq1Γ(q1){n=0q12(1)q12+nαn,qi=02n1(2n1i)(i)2n1iω2n1i[sζH(i,iω)+(1)i+1sζH(i,iω)]q:odd,n=0q21(1)q21+nβn,qi=02n(2ni)(i)2niω2ni[sζH(i,12+iω)+(1)isζH(i,12iω)]q:even.\displaystyle\begin{aligned} \partial_{s}\zeta_{\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}}^{(2)}(0)&=2c_{p+1}\,r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\frac{2r_{q-1}}{\Gamma(q-1)}\\ &\qquad\cdot\begin{dcases}\begin{aligned} &\sum_{n=0}^{\frac{q-1}{2}}(-1)^{\frac{q-1}{2}+n}\,\alpha_{n,q}\,\sum_{i=0}^{2n-1}\binom{2n-1}{i}(-\textrm{i})^{2n-1-i}\omega^{2n-1-i}\\ &\quad\cdot\left[\partial_{s}\zeta_{\text{H}}\left(-i,\textrm{i}\,\omega\right)+(-1)^{i+1}\partial_{s}\zeta_{\text{H}}\left(-i,-\textrm{i}\,\omega\right)\right]\end{aligned}&q:\text{odd}\,,\\ \begin{aligned} &\sum_{n=0}^{\frac{q}{2}-1}(-1)^{\frac{q}{2}-1+n}\,\beta_{n,q}\,\sum_{i=0}^{2n}\binom{2n}{i}(-\textrm{i})^{2n-i}\omega^{2n-i}\\ &\quad\cdot\left[\partial_{s}\zeta_{\text{H}}\left(-i,\frac{1}{2}+\textrm{i}\,\omega\right)+(-1)^{i}\partial_{s}\zeta_{\text{H}}\left(-i,\frac{1}{2}-\textrm{i}\,\omega\right)\right]\end{aligned}&q:\text{even}\,.\end{dcases}\end{aligned} (124)

For even qq, the combination of the Hurwitz zeta functions can be simplified using a formula

sζH(i,12+iω)+(1)isζH(i,12iω)=Γ(i+1)(2πi)iLii+1(e2πω)+πii+1Bi+1(12+iω).\displaystyle\begin{aligned} &\partial_{s}\zeta_{\text{H}}\left(-i,\frac{1}{2}+\textrm{i}\,\omega\right)+(-1)^{i}\partial_{s}\zeta_{\text{H}}\left(-i,\frac{1}{2}-\textrm{i}\,\omega\right)\\ &\qquad=\frac{\Gamma(i+1)}{(2\pi\textrm{i})^{i}}\mathrm{Li}_{i+1}(-\mathrm{e}^{-2\pi\omega})+\frac{\pi\,\textrm{i}}{i+1}B_{i+1}\left(\frac{1}{2}+\textrm{i}\,\omega\right)\,.\end{aligned} (125)

See e.g. (B.19) in Nishioka:2021uef for the derivation. Unfortunately, it is difficult to simplify the equations anymore, so we perform the integral (124) numerically.

For odd pp, we find the following:

  • For odd qq, we numerically find the relation among free energies

    Fren[𝕊d]=Fren[2k×𝕊d2k]\displaystyle F_{\text{ren}}[\mathbb{S}^{d}]=F_{\text{ren}}[\mathbb{H}^{2k}\times\mathbb{S}^{d-2k}] (126)

    for k=1,2,,d/21k=1,2,\cdots,d/2-1. That is, the anomaly coefficients and the finite parts satisfy the relations,

    A[𝕊d]\displaystyle A[\mathbb{S}^{d}] =A[2k×𝕊d2k],\displaystyle=A[\mathbb{H}^{2k}\times\mathbb{S}^{d-2k}]\,, (127)
    Ffin[𝕊d]\displaystyle F_{\text{fin}}[\mathbb{S}^{d}] =Ffin[2k×𝕊d2k].\displaystyle=F_{\text{fin}}[\mathbb{H}^{2k}\times\mathbb{S}^{d-2k}]\,. (128)
  • For even qq, the anomaly parts vanish since the bulk dimension d=p+qd=p+q is odd. We numerically find the relation between the universal parts of the free energy

    Ffin[𝕊d]=Ffin[2k×𝕊d2k]\displaystyle F_{\text{fin}}[\mathbb{S}^{d}]=F_{\text{fin}}[\mathbb{H}^{2k}\times\mathbb{S}^{d-2k}] (129)

    for k=1,2,,(d1)/2k=1,2,\cdots,(d-1)/2.

The equivalence of the anomaly coefficients between 2k×𝕊d2k\mathbb{H}^{2k}\times\mathbb{S}^{d-2k} and 𝕊d\mathbb{S}^{d} follows from a relation of Euler characteristic χ[2k×𝕊d2k]=χ[𝕊d]\chi[\mathbb{H}^{2k}\times\mathbb{S}^{d-2k}]=\chi[\mathbb{S}^{d}] as pointed out Rodriguez-Gomez:2017kxf ; Nishioka:2021uef because the bulk anomaly is related to the Euler characteristic.

5 Free energy for Neumann boundary condition

In section 4, we computed the free energy on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} with Dirichlet boundary condition. In this section, we compute the free energy with Neumann boundary condition. Neumann boundary condition exists only when q=2q=2 as we saw in section 2.

5.1 Analytical continuation

We decompose the free energy into a sum of \ell,

F[p+1×𝕊q1]==0g(q1)()F(ν(q1)),\displaystyle F[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]=\sum_{\ell=0}^{\infty}g^{(q-1)}(\ell)F_{\ell}\left(\nu_{\ell}^{(q-1)}\right)\,, (130)

where F(ν(q1))F_{\ell}\left(\nu_{\ell}^{(q-1)}\right) is the free energy for the \ell-th mode on p+1\mathbb{H}^{p+1},

F(ν(q1))=dU(1)20dωμ(p+1)(ω)[log(ω+iν(q1)Λ~R)+log(ωiν(q1)Λ~R)].\displaystyle F_{\ell}\left(\nu_{\ell}^{(q-1)}\right)=-\frac{d_{U(1)}}{2}\int_{0}^{\infty}\mathrm{d}\omega\,\mu^{(p+1)}(\omega)\,\left[\log\left(\frac{\omega+\textrm{i}\,\nu_{\ell}^{(q-1)}}{\tilde{\Lambda}R}\right)+\log\left(\frac{\omega-\textrm{i}\,\nu_{\ell}^{(q-1)}}{\tilde{\Lambda}R}\right)\right]\,. (131)

Since the ω\omega integral is performed before the summation over \ell, the logarithmic function is decomposed in this way.

From now on, we concentrate on q=2q=2 because Neumann boundary condition is allowed only when q=2q=2. The difference of free energies between the two boundary conditions comes from the =0\ell=0 mode. The Dirichlet boundary condition has a positive value

ν=0(1)=12,\displaystyle\nu_{\ell=0}^{(1)}=\frac{1}{2}\,, (132)

while the Neumann boundary condition has a negative value

ν=0(1)=12.\displaystyle\nu_{\ell=0}^{(1)}=-\frac{1}{2}\,. (133)

In section 4.1, we obtained the zeta functions as analytical functions of positive mm. It is possible to analytically continue the zeta functions to a m<0m<0 region.

Even pp

From (80) the free energy of the =0\ell=0 mode is given by

F=0(ν=0(1))=(1)p2+12p2Γ(p+1)k=0p2βk,p+2(ν=0(1))2k+12k+1(1)klog(Rϵ).\displaystyle F_{\ell=0}\left(\nu_{\ell=0}^{(1)}\right)=\frac{(-1)^{\frac{p}{2}+1}2^{\frac{p}{2}}}{\Gamma(p+1)}\sum_{k=0}^{\frac{p}{2}}\beta_{k,p+2}\frac{\left(\nu_{\ell=0}^{(1)}\right)^{2k+1}}{2k+1}(-1)^{k}\log\left(\frac{R}{\epsilon}\right)\,. (134)

Then, the difference of the free energies becomes

FΔD[p+1×𝕊1]FΔN[p+1×𝕊1]=(2)p2+1Γ(p+1)k=0p2βk,p+2(1)k22k(2k+1)log(Rϵ).\displaystyle F_{\Delta_{\text{D}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]-F_{\Delta_{\text{N}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]=\frac{(-2)^{\frac{p}{2}+1}}{\Gamma(p+1)}\sum_{k=0}^{\frac{p}{2}}\beta_{k,p+2}\frac{(-1)^{k}}{2^{2k}(2k+1)}\log\left(\frac{R}{\epsilon}\right)\,. (135)

By using

122k(2k+1)=2012duu2k,\displaystyle\frac{1}{2^{2k}(2k+1)}=2\int_{0}^{\frac{1}{2}}\!\mathrm{d}u\,u^{2k}\,, (136)

(45) with the replacement ν(d)u\nu_{\ell}^{(d)}\to u, and the reflection formula of the gamma function

Γ(z)Γ(1z)=πsin(πz)\displaystyle\Gamma(z)\Gamma(1-z)=\frac{\pi}{\sin(\pi z)} (137)

with z=up+12+1z=u-\frac{p+1}{2}+1, the coefficient of the logarithmic divergent part becomes

(2)p2+1Γ(p+1)k=0p2βk,p+2(1)k22k(2k+1)=2(2)p2+1πΓ(p+1)012ducos(πu)Γ(p+12+u)Γ(p+12u).\displaystyle\frac{(-2)^{\frac{p}{2}+1}}{\Gamma(p+1)}\sum_{k=0}^{\frac{p}{2}}\beta_{k,p+2}\frac{(-1)^{k}}{2^{2k}(2k+1)}=\frac{2(-2)^{\frac{p}{2}+1}}{\pi\Gamma(p+1)}\int_{0}^{\frac{1}{2}}\!\mathrm{d}u\,\cos(\pi u)\Gamma\left(\frac{p+1}{2}+u\right)\Gamma\left(\frac{p+1}{2}-u\right)\,. (138)

We find a relation

FΔD[p+1×𝕊1]FΔN[p+1×𝕊1]=2Funiv[𝕊p],\displaystyle F_{\Delta_{\text{D}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]-F_{\Delta_{\text{N}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]=-2F_{\text{univ}}[\mathbb{S}^{p}]\,, (139)

where the universal free energy on 𝕊p\mathbb{S}^{p} is given by (58).

Odd pp

From (94) and (95) with (152), the difference of the free energies is given by

FΔD[p+1×𝕊1]FΔN[p+1×𝕊1]=2cp+1rp+1k=1p+12αk,p+2(fk(12)fk(12))=2cp+1rp+1k=1p+12αk,p+2(1)k(122k1(2k1)+1212duu2k1ψ(u)).\displaystyle\begin{aligned} &F_{\Delta_{\text{D}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]-F_{\Delta_{\text{N}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]\\ &\qquad=-2c_{p+1}\,r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}\left(f_{k}\left(\frac{1}{2}\right)-f_{k}\left(-\frac{1}{2}\right)\right)\\ &\qquad=-2c_{p+1}\,r_{p+1}\sum_{k=1}^{\frac{p+1}{2}}\alpha_{k,p+2}(-1)^{k}\left(\frac{1}{2^{2k-1}(2k-1)}+\int_{-\frac{1}{2}}^{\frac{1}{2}}\!\mathrm{d}u\,u^{2k-1}\psi(u)\right)\,.\end{aligned} (140)

By using the identity

ψ(u)ψ(u)=1u+πcot(πu),\displaystyle\psi(u)-\psi(-u)=\frac{1}{u}+\pi\cot(\pi u)\,, (141)

(45) with the replacement ν(d)u\nu_{\ell}^{(d)}\to u and the reflection formula of the gamma function with z=up+12+1z=u-\frac{p+1}{2}+1, we find

FΔD[p+1×𝕊1]FΔN[p+1×𝕊1]=2cp+1rp+1012ducos(πu)Γ(p+12+u)Γ(p+12u)=2Funiv[𝕊p],\displaystyle\begin{aligned} F_{\Delta_{\text{D}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]-F_{\Delta_{\text{N}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]&=2c_{p+1}r_{p+1}\int_{0}^{\frac{1}{2}}\!\mathrm{d}u\,\cos(\pi u)\Gamma\left(\frac{p+1}{2}+u\right)\Gamma\left(\frac{p+1}{2}-u\right)\\ &=-2F_{\text{univ}}[\mathbb{S}^{p}]\,,\end{aligned} (142)

where the universal free energy on 𝕊p\mathbb{S}^{p} is given by (58).

Independent of the evenness of pp, the difference of the free energies between the two boundary conditions is proportional to the universal free energy of the fermion on 𝕊p\mathbb{S}^{p},

FΔD[p+1×𝕊1]FΔN[p+1×𝕊1]=2Funiv[𝕊p].\displaystyle F_{\Delta_{\text{D}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]-F_{\Delta_{\text{N}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]=-2F_{\text{univ}}[\mathbb{S}^{p}]\,. (143)

This fact implies that the Neumann boundary condition for q=2q=2 is trivial in the sense that the defect operator saturates the unitarity bound (31) and becomes a free field.

5.2 Evidence for defect CC-theorem

We are now in a position to compare the result in section 5.1 with our proposed conjecture in Kobayashi:2018lil . The defect free energy (5) may not be invariant under the Weyl transformation, while we expect that the difference of the defect free energies is invariant. An RG flow from the Neumann boundary condition to the Dirichlet boundary condition is triggered by a double trace deformation as is familiar in the AdS/CFT setup Witten:2001ua ; Berkooz:2002ug ; Gubser:2002zh ; Gubser:2002vv ; Hartman:2006dy ; Diaz:2007an ; Giombi:2013yva , and the difference of the universal part of the defect free energies is given by

D~UVD~IR=sin(πp2)(FΔN[p+1×𝕊1]FΔD[p+1×𝕊1])=2F~[𝕊p],\displaystyle\begin{aligned} \tilde{D}_{\text{UV}}-\tilde{D}_{\text{IR}}&=-\sin\left(\frac{\pi p}{2}\right)\left(F_{\Delta_{\text{N}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]-F_{\Delta_{\text{D}}}[\mathbb{H}^{p+1}\times\mathbb{S}^{1}]\right)\\ &=2\tilde{F}[\mathbb{S}^{p}]\,,\end{aligned} (144)

where

F~[𝕊p]=2rpΓ(p+1)012ducos(πu)Γ(p+12+u)Γ(p+12u)\displaystyle\tilde{F}[\mathbb{S}^{p}]=\frac{2r_{p}}{\Gamma(p+1)}\int_{0}^{\frac{1}{2}}\!\mathrm{d}u\,\cos(\pi u)\Gamma\left(\frac{p+1}{2}+u\right)\Gamma\left(\frac{p+1}{2}-u\right) (145)

is positive for any pp. The positivity of the sphere free energy leads to the positivity of the difference of the free energies between at UV fixed point and at IR fixed point. In this case, our proposed defect CC-theorem holds.

6 Summary and Discussion

In this paper, we studied a free Dirac fermion on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} as a DCFT. In section 2, we classified the allowed boundary conditions, and we found that a nontrivial boundary condition is allowed only in q=2q=2. In sections 3 and 4, we computed the free energy on 𝕊d\mathbb{S}^{d}, 𝕊d\mathbb{HS}^{d}, d\mathbb{H}^{d} and p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} with the Dirichlet boundary condition using the zeta-function regularisation. In particular, we obtained relations of free energies, which hold also in a conformally coupled scalar field Rodriguez-Gomez:2017kxf ; Nishioka:2021uef in section 4.2. In section 5, we computed the difference of the free energies between the Dirichlet boundary condition and the Neumann boundary condition in q=2q=2 and confirmed the validity of our proposed defect CC-theorem.

We obtained various results similar to a conformally coupled scalar case Rodriguez-Gomez:2017kxf ; Nishioka:2021uef . However, there are several differences between the fermion case and the scalar case. The first difference comes from the codimension of the defect which allows nontrivial boundary conditions. Nontrivial boundary conditions in the conformally coupled scalar are allowed in q=1,2,3,4q=1,2,3,4 Nishioka:2021uef , but the nontrivial boundary condition in the free fermion occurs in q=2q=2. The second difference is that we rigorously derive the equivalence of the free energies between 𝕊d\mathbb{HS}^{d} and d\mathbb{H}^{d} for arbitrary dd. However, for the conformally coupled scalar Nishioka:2021uef , the equivalence of the free energies between 𝕊d\mathbb{HS}^{d} and d\mathbb{H}^{d} is checked only numerically because a nontrivial identity among Bernoulli polynomials are required for a proof of the equivalence for arbitrary dd.

In section 2.2, we gave a classification of boundary conditions in a free fermion, and this means that we constructed a concrete model of a DCFT. A task to derive the same classification (or defect operator) of boundary conditions by using the method in Lauria:2020emq remains. The concrete model would be useful for a study of DCFTs. In this paper we discussed the non-monodromy defect where the fermion does not receive any phase around the defect. For q=2q=2, there exits a monodromy defect where the fermion receives a phase around the defect in addition to the non-monodromy defect.

Acknowledgements.
The author thanks T. Nishioka for the collaboration of the related work, useful discussion and various comments on the draft of this paper. The author thanks D. Rodriguez-Gomez and J. G. Russo for comments on the draft of this paper and useful discussions. The author also thanks Y. Abe, Y. Okuyama and M. Watanabe for useful communication. The work is supported by the National Center of Theoretical Sciences (NCTS).

Appendix A List of tables

\cdot 𝕊1\mathbb{S}^{1} 𝕊2\mathbb{S}^{2} 𝕊3\mathbb{S}^{3} 𝕊4\mathbb{S}^{4}
\cdot 0 13\frac{1}{3} 0 1190-\frac{11}{90}
2\mathbb{H}^{2} 16\frac{1}{6} 0 1190-\frac{11}{90} 0 1913780\frac{191}{3780}
3\mathbb{H}^{3} 0 371440-\frac{37}{1440} 0 1917560\frac{191}{7560} 0
4\mathbb{H}^{4} 11180-\frac{11}{180} 0 1913780\frac{191}{3780} 0 2497113400-\frac{2497}{113400}
5\mathbb{H}^{5} 0 40740320\frac{407}{40320} 0 15195114515200-\frac{151951}{14515200} 0
6\mathbb{H}^{6} 1917560\frac{191}{7560} 0 2497113400-\frac{2497}{113400} 0 147971496880\frac{14797}{1496880}
7\mathbb{H}^{7} 0 12460329030400-\frac{124603}{29030400} 0 4384643958003200\frac{4384643}{958003200} 0
8\mathbb{H}^{8} 2497226800-\frac{2497}{226800} 0 147971496880\frac{14797}{1496880} 0 9242715720432412000-\frac{92427157}{20432412000}
9\mathbb{H}^{9} 0 72779333832012800\frac{7277933}{3832012800} 0 4310521477320922789888000-\frac{43105214773}{20922789888000} 0
10\mathbb{H}^{10} 147972993760\frac{14797}{2993760} 0 9242715720432412000-\frac{92427157}{20432412000} 0 3674061717513496000\frac{36740617}{17513496000}
𝕊5\mathbb{S}^{5} 𝕊6\mathbb{S}^{6} 𝕊7\mathbb{S}^{7} 𝕊8\mathbb{S}^{8}
\cdot 0 1913780\frac{191}{3780} 0 2497113400-\frac{2497}{113400}
2\mathbb{H}^{2} 0 2497113400-\frac{2497}{113400} 0 147971496880\frac{14797}{1496880}
3\mathbb{H}^{3} 335332903040-\frac{33533}{2903040} 0 726491136857600\frac{726491}{136857600} 0
4\mathbb{H}^{4} 0 147971496880\frac{14797}{1496880} 0 9242715720432412000-\frac{92427157}{20432412000}
5\mathbb{H}^{5} 147972993760\frac{14797}{2993760} 0 34676277671494484992000-\frac{3467627767}{1494484992000} 0
6\mathbb{H}^{6} 0 9242715720432412000-\frac{92427157}{20432412000} 0 3674061717513496000\frac{36740617}{17513496000}
7\mathbb{H}^{7} 46098620032092278988800-\frac{4609862003}{2092278988800} 0 3674061735026992000\frac{36740617}{35026992000} 0
8\mathbb{H}^{8} 0 3674061717513496000\frac{36740617}{17513496000} 0 6143094316962523180720000-\frac{61430943169}{62523180720000}
9\mathbb{H}^{9} 5045369643750214695731200\frac{50453696437}{50214695731200} 0 2020070414498341811420119040000-\frac{20200704144983}{41811420119040000} 0
10\mathbb{H}^{10} 0 6143094316962523180720000-\frac{61430943169}{62523180720000} 0 2313394589230349893498214560000\frac{23133945892303}{49893498214560000}
Table 2: The bulk anomalies A[p+1×𝕊q1]A[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}] and the defect anomalies 𝒜[p+1×𝕊q1]\mathcal{A}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}] (shaded) on p+1×𝕊q1\mathbb{H}^{p+1}\times\mathbb{S}^{q-1} with the Dirichlet boundary condition.
\mathcal{M} Ffin[]F_{\text{fin}}[\mathcal{M}]
𝕊2\mathbb{S}^{2} 4ζ(1)4\,\zeta^{\prime}(-1)
𝕊3\mathbb{S}^{3} 14log2+38π2ζ(3)\frac{1}{4}\,\log 2+\frac{3}{8\pi^{2}}\,\zeta(3)
𝕊4\mathbb{S}^{4} 43ζ(1)+43ζ(3)-\frac{4}{3}\,\zeta^{\prime}(-1)+\frac{4}{3}\,\zeta^{\prime}(-3)
𝕊5\mathbb{S}^{5} 332log2532π2ζ(3)1564π4ζ(5)-\frac{3}{32}\,\log 2-\frac{5}{32\pi^{2}}\,\zeta(3)-\frac{15}{64\pi^{4}}\,\zeta(5)
𝕊6\mathbb{S}^{6} 815ζ(1)23ζ(3)+215ζ(5)\frac{8}{15}\,\zeta^{\prime}(-1)-\frac{2}{3}\,\zeta^{\prime}(-3)+\frac{2}{15}\,\zeta^{\prime}(-5)
𝕊7\mathbb{S}^{7} 5128log2+2593840π2ζ(3)+35256π4ζ(5)+63512π6ζ(7)\frac{5}{128}\,\log 2+\frac{259}{3840\pi^{2}}\,\zeta(3)+\frac{35}{256\pi^{4}}\,\zeta(5)+\frac{63}{512\pi^{6}}\,\zeta(7)
𝕊8\mathbb{S}^{8} 835ζ(1)+1445ζ(3)445ζ(5)+2315ζ(7)-\frac{8}{35}\,\zeta^{\prime}(-1)+\frac{14}{45}\,\zeta^{\prime}(-3)-\frac{4}{45}\,\zeta^{\prime}(-5)+\frac{2}{315}\,\zeta^{\prime}(-7)
𝕊9\mathbb{S}^{9} 352048log23229107520π2ζ(3)1412048π4ζ(5)1892048π6ζ(7)2554096π8ζ(9)-\frac{35}{2048}\,\log 2-\frac{3229}{107520\pi^{2}}\,\zeta(3)-\frac{141}{2048\pi^{4}}\,\zeta(5)-\frac{189}{2048\pi^{6}}\,\zeta(7)-\frac{255}{4096\pi^{8}}\,\zeta(9)
𝕊10\mathbb{S}^{10} 32315ζ(1)82567ζ(3)+13270ζ(5)1189ζ(7)+15670ζ(9)\frac{32}{315}\,\zeta^{\prime}(-1)-\frac{82}{567}\,\zeta^{\prime}(-3)+\frac{13}{270}\,\zeta^{\prime}(-5)-\frac{1}{189}\,\zeta^{\prime}(-7)+\frac{1}{5670}\,\zeta^{\prime}(-9)
Odd dd Ffin[𝕊d]=12Ffin[𝕊d]F_{\text{fin}}[\mathbb{HS}^{d}]=\frac{1}{2}F_{\text{fin}}[\mathbb{S}^{d}]
Even dd Ffin[𝕊d]=12Ffin[𝕊d]F_{\text{fin}}[\mathbb{HS}^{d}]=\frac{1}{2}F_{\text{fin}}[\mathbb{S}^{d}]
Odd dd 0
Even dd Ffin[d]=12Ffin[𝕊d]F_{\text{fin}}[\mathbb{H}^{d}]=\frac{1}{2}F_{\text{fin}}[\mathbb{S}^{d}]
Even pp Ffin[p+1×𝕊q1]=0F_{\text{fin}}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}]=0
Odd pp Ffin[𝕊d]=Ffin[2k×𝕊d2k]F_{\text{fin}}[\mathbb{S}^{d}]=F_{\text{fin}}[\mathbb{H}^{2k}\times\mathbb{S}^{d-2k}] for k=1,,d/21k=1,\cdots,\lceil d/2\rceil-1
Table 3: Table of the finite parts of Ffin[𝕊d]F_{\text{fin}}[\mathbb{S}^{d}], Ffin[𝕊d]F_{\text{fin}}[\mathbb{HS}^{d}], Ffin[d]F_{\text{fin}}[\mathbb{H}^{d}], and Ffin[p+1×𝕊q1]F_{\text{fin}}[\mathbb{H}^{p+1}\times\mathbb{S}^{q-1}] with the Dirichlet boundary condition.

Appendix B Detail derivation of (92)

In this appendix, we give a detailed derivation of (92).

To perform the integral

fk(m)=0dωω2k1e2πω1log(ω2+m2),\displaystyle f_{k}(m)=\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\log(\omega^{2}+m^{2})\,, (146)

we first take the derivative respect to mm,

mfk(m)\displaystyle\partial_{m}f_{k}(m) =2mgk(m),\displaystyle=2mg_{k}(m)\,, (147)
gk(m)\displaystyle g_{k}(m) =0dωω2k1(e2πω1)(ω2+m2).\displaystyle=\int_{0}^{\infty}\!\mathrm{d}\omega\,\frac{\omega^{2k-1}}{(\mathrm{e}^{2\pi\omega}-1)(\omega^{2}+m^{2})}\,. (148)

Since gk(m)g_{k}(m) satisfies the recursion relation

gk+1(m)=m2gk(m)+(1)k+1B2k4k,\displaystyle g_{k+1}(m)=-m^{2}g_{k}(m)+(-1)^{k+1}\frac{B_{2k}}{4k}\,, (149)

with the initial condition

g1(m)=12(logm12mψ(m)),\displaystyle g_{1}(m)=\frac{1}{2}\left(\log m-\frac{1}{2m}-\psi(m)\right)\,, (150)

the solution is given by

gk(m)=(1)k1m2k2(g1(m)l=1k1m2lB2l4l).\displaystyle g_{k}(m)=(-1)^{k-1}m^{2k-2}\left(g_{1}(m)-\sum_{l=1}^{k-1}m^{-2l}\frac{B_{2l}}{4l}\right)\,. (151)

By integrating 2mgk(m)2mg_{k}(m) from 0 to mm, fk(m)f_{k}(m) can be evaluated as

fk(m)=(1)k[ζ(12k)+m2k12(2k1)+m2k4k(1klog(m2))+0mdμμ2k1ψ(μ)+l=1k1B2l4lm2k2lkl],\displaystyle\begin{aligned} f_{k}(m)&=(-1)^{k}\left[-\zeta^{\prime}(1-2k)+\frac{m^{2k-1}}{2(2k-1)}+\frac{m^{2k}}{4k}\left(\frac{1}{k}-\log(m^{2})\right)\right.\\ &\qquad\qquad\left.+\int_{0}^{m}\!\mathrm{d}\mu\,\mu^{2k-1}\psi(\mu)+\sum_{l=1}^{k-1}\frac{B_{2l}}{4l}\frac{m^{2k-2l}}{k-l}\right]\,,\end{aligned} (152)

where we use

fk(0)=20dωω2k1e2πω1logω=(1)k1ζ(12k).\displaystyle\begin{aligned} f_{k}(0)&=2\int_{0}^{\infty}\!\mathrm{d}\omega\frac{\omega^{2k-1}}{\mathrm{e}^{2\pi\omega}-1}\log\omega\\ &=(-1)^{k-1}\zeta^{\prime}(1-2k)\,.\end{aligned} (153)

The remaining integral is performed using a formula in adamchik1998polygamma ; espinosa2001some

0mdμμnψ(μ)=(1)n(Bn+1Hnn+1ζ(n))+r=0n(1)r(nr)mnr(ζ(r,m)Br+1(m)Hrr+1).\displaystyle\begin{aligned} \int_{0}^{m}\!\mathrm{d}\mu\,\mu^{n}\psi(\mu)&=(-1)^{n}\left(\frac{B_{n+1}H_{n}}{n+1}-\zeta^{\prime}(-n)\right)\\ &\qquad+\sum_{r=0}^{n}(-1)^{r}\binom{n}{r}m^{n-r}\left(\zeta^{\prime}(-r,m)-\frac{B_{r+1}(m)H_{r}}{r+1}\right)\,.\end{aligned} (154)

References