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

institutetext: Department of Mathematics, King’s College London, Strand, London, WC2R 2LS, UK

Effective theory for fusion of conformal defects

Petr Kravchuk, Alex Radcliffe, Ritam Sinha
Abstract

We construct an effective field theory for fusion of conformal defects of any codimension in d3d\geq 3 conformal field theories. We fully solve the constraints of Weyl invariance for defects of arbitrary shape on general curved bulk manifolds and discuss the simplifications that arise for spherical defects on the conformal sphere. As applications, we study the structure of cusp anomalous dimensions in the anti-parallel lines limit and derive high-energy spin-dependent asymptotics for the one-point functions of bulk operators. We point out the potential importance of defects that break transverse rotations and initiate a classification of their Weyl anomalies.

1 Introduction

In this paper we study the limit of CFT correlation functions in which two conformal defects are brought close together, i.e. the problem of “fusion” of conformal defects. While the analogous problem for topological defects has a long history, the conformal case has received much less attention. In d=2d=2 dimensions, it has been studied in Bachas:2007td ; Bachas:2013ora ; Konechny:2015qla . In d3d\geq 3, fusion of conformal defects (or correlation functions of two defects) was analyzed in Soderberg:2021kne ; Rodriguez-Gomez:2022gbz ; SoderbergRousu:2023zyj .111Furthermore, the papers Diatlyk:2024qpr ; Diatlyk:2024zkk appeared while this paper was in preparation, see the end of this section for further comments. For the most part, these works focused on calculations in specific theories. The main goal of this work is to study the general properties of fusion of conformal defects and to explore some initial applications.

To be more concrete, we consider a correlation function

𝒟1𝒟2\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\cdots\rangle (1)

with insertions of two conformal defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. In this paper, we use \langle\cdots\rangle to denote partition functions normalised so that the partition function with no insertions is 11. In particular, we do not normalise by defect partition functions in any way.

For simplicity, we assume that no defect local operators are present on 𝒟i\mathcal{D}_{i}, although the forthcoming discussion can be easily adapted to the more general case. The dots “\cdots” represent any other insertions which stay O(1)O(1) distance away from the defects 𝒟i\mathcal{D}_{i}. We do not make any assumptions about the shape or the topology of the defects 𝒟i\mathcal{D}_{i}, except that they do not intersect and are homotopic to each other, see figure 1. Our goal is to determine the behaviour of the correlation function (1) in the limit where 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} approach each other and eventually fully overlap. In other words, if LL is the typical separation between 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}, we are interested in the limit L0L\to 0.

Our main assumption is that the appropriate language for this problem is that of a low-energy effective theory (EFT) on top of a new conformal defect 𝒟Σ\mathcal{D}_{\Sigma},

𝒟1𝒟2𝒟Σ[eSeff],\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\cdots\rangle\sim\langle\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}}]\cdots\rangle, (2)

where SeffS_{\text{eff}} describes the perturbation of the conformal defect 𝒟Σ\mathcal{D}_{\Sigma} by identity and the irrelevant operators. The effective action SeffS_{\text{eff}} depends on the relative configuration of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}, and the high-dimension operators are suppressed by powers of defect separation LL. In practice, this EFT will express the correlation function (1) as an asymptotic series in powers of L/RL/R, where RR is the IR length scale set by the size of 𝒟Σ\mathcal{D}_{\Sigma} and the other insertions.

Refer to caption𝒟1\mathcal{D}_{1}𝒟2\mathcal{D}_{2}LL
Refer to caption
Figure 1: Examples of fusion configurations. Left: a simple fusion configuration, the separation scale LL is taken to 0. Right: a generic fusion configuration with knotted and interlinked defects.

We defer the more precise discussion of (2) to section 3, and discuss here only the motivation behind and the status of this assumption. The key intuition is that for small LL, we may view the pair of defects 𝒟1,𝒟2\mathcal{D}_{1},\mathcal{D}_{2} as giving a UV definition of a new non-conformal defect. In this picture LL plays the role of the UV scale analogous to, for example, a lattice scale. The process of taking the limit L0L\to 0 (equivalently, RR\to\infty) can be then understood as an RG flow, which we expect to generically terminate at a conformal fixed point 𝒟Σ\mathcal{D}_{\Sigma}. Small but finite LRL\ll R can then be described in terms of an EFT around 𝒟Σ\mathcal{D}_{\Sigma}.222The defect 𝒟Σ\mathcal{D}_{\Sigma} need not be simple, although we expect that generically it will be, see section 3.5. Note that this point of view on fusion is not by any means new and was discussed already in Bachas:2007td .

Our main result is a general formalism for writing down the effective action SeffS_{\text{eff}} appearing in (2), properly taking into account the constraints imposed by Weyl invariance on either side of the equation. We describe it in section 3. In the case of local operator product expansion, Weyl invariance fixes the form of the contribution of descendants in terms of that of the primary. In the defect case, it constrains the terms allowed in the effective action, and we classify all possible terms. In order to achieve this, we derive in section 2 a Weyl-invariant parameterisation of the relative position of the two defects and show that it is possible to construct from it a Weyl-invariant bulk metric g^\widehat{g}, to which we refer as the fusion metric.

As an application of the general formalism, we study in section 4 the implications of the effective theory for the cusp anomalous dimension Γcusp\Gamma_{\text{cusp}}, which appears when two line defects meet with an opening angle α\alpha. We show that a systematic small-α\alpha expansion of Γcusp\Gamma_{\text{cusp}} and its excited versions can be computed from SeffS_{\text{eff}}. We discuss some simple consequences of this for supersymmetric Wilson lines in 𝒩=4\mathcal{N}=4 SYM and compute a subleading Wilson coefficient from the known expressions for Γcusp\Gamma_{\text{cusp}}. We also make some comments regarding the connection of “ultra-soft” degrees of freedom Pineda:2007kz ; Correa:2012nk and the simplicity of 𝒟Σ\mathcal{D}_{\Sigma} in section 3.5.

As another application, in section 5 we consider the two-point function 𝒟𝒟¯\langle\mathcal{D}\overline{\mathcal{D}}\rangle of a defect and its conjugate from two points of view. On the one hand, the asymptotic expansion (2) applies. On the other hand, each defect can be expanded in terms of local operators Gadde:2016fbj . Consistency of these two expansions allows us to derive high-energy asymptotics of the one-point functions 𝒪𝒟\langle{\mathcal{O}}\mathcal{D}\rangle of local operators 𝒪{\mathcal{O}} in the presence of the defect 𝒟\mathcal{D}, as functions of both the scaling dimension and the spin of 𝒪{\mathcal{O}}.333A similar computation of the density of states in terms of thermal EFT was recently performed in Benjamin:2023qsc .

An important contribution to the effective action SeffS_{\text{eff}} comes from terms which compensate for the difference of Weyl anomalies between the two sides of (2). This is however complicated by the fact that 𝒟Σ\mathcal{D}_{\Sigma} does not have to preserve the same space-time symmetries as 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. Indeed, if we view the pair 𝒟1𝒟2\mathcal{D}_{1}\mathcal{D}_{2} as the UV definition of an RG flow, this definition explicitly breaks transverse rotations around the defects. This transverse rotation symmetry may or may not be restored at the IR fixed point 𝒟Σ\mathcal{D}_{\Sigma}. If it is not, we show in section 6 that the possible Weyl anomalies of 𝒟Σ\mathcal{D}_{\Sigma} are more general than usually considered in the literature (see Chalabi:2021jud for a review), and classify them in the case of line defects in general dd and surface defects in d=4d=4.

While we believe that the argument in favor of (2) presented above is generically correct, we stress that it is not something that we can prove (with a reasonable degree of rigor) in general CFTs. This is contrast to the situation with the local operator product expansion (OPE), where quite formal arguments exist Mack:1976pa . Even though local OPE can be formally viewed as fusion of 0-dimensional defects, one has to be cautious about trying to take this analogy too far. We discuss the similarities and the key differences between fusion of conformal defects and local OPE in section 7.

In section 1.1 we give a somewhat more detailed summary of our findings.

We conclude in section 8. Appendices contain some additional details, most notably the 2-derivative couplings for the identity operator are classified in appendix C.

Note added

When this work was largely complete, the papers Diatlyk:2024zkk ; Diatlyk:2024qpr appeared which partially overlap with our discussion. The paper Diatlyk:2024zkk discussed some general properties of defect fusion but did not study the effective action beyond the leading cosmological constant term, focusing instead on multiple explicit examples of fusion. We have added comparisons of some our results with these examples where applicable. The more recent paper Diatlyk:2024qpr studied fusion in the case when 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are conformal boundaries with the bulk CFT in between them. They also focused only on the leading term and derived the high-energy asymptotics of one-point functions. This partially overlaps with the more general results of our section 5. We have also learned from Diatlyk:2024qpr of an upcoming related work zoharfuture .

1.1 Summary

This section contains a condensed summary of our main results. For the more detailed discussion we refer the reader to the main body of the paper.

We work in a conformal field theory in d>2d>2 dimensions and consider a pair of pp-dimensional conformal defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} which are approximately parallel to one another, see figure 1. We denote the codimension by q=dpq=d-p. We are interested in the limit in which the two defects fuse. That is, we consider a smooth family of configurations parameterised by an “average distance” L>0L>0 such that in the limit L0L\to 0 the two defects coincide. We only use LL to denote a scale (except in some examples where it is given a precise meaning), and the precise dependence on the shape of the defects will be captured by other data.

At fixed LL, a generic configuration of this sort breaks most conformal symmetries even if 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are spherical defects in flat space. We work under the assumption that in the limit L0L\to 0 the conformal symmetry is restored, and a conformal defect 𝒟Σ\mathcal{D}_{\Sigma} is obtained. Such a fusion process can be described in terms of an RG flow. In this picture, the product 𝒟1𝒟2\mathcal{D}_{1}\mathcal{D}_{2} can be viewed as the UV definition of a non-conformal defect which in the IR flows to 𝒟Σ\mathcal{D}_{\Sigma}. In the neighbourhood of the IR fixed point 𝒟Σ\mathcal{D}_{\Sigma}, i.e. in the limit of small LL, this RG flow can be described in terms of an effective action SeffS_{\text{eff}} on top of 𝒟Σ\mathcal{D}_{\Sigma}. Our goal is to analyze constraints on 𝒟Σ\mathcal{D}_{\Sigma} and to describe the general form of the effective action SeffS_{\text{eff}}.

Kinematics

To study these questions systematically, we consider the problem on a general curved bulk manifold MM and allow arbitrarily shaped 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. We assume that the bulk theory and all conformal defects are diffeomorphism- and Weyl-invariant (modulo Weyl anomalies). This allows us to use diffeomorphism and Weyl invariance to constrain 𝒟Σ\mathcal{D}_{\Sigma} and the effective action SeffS_{\text{eff}}.

The effective action SeffS_{\text{eff}} has to depend on the relative position of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. To the leading order in LL we can parameterise the position of 𝒟2\mathcal{D}_{2} relative to 𝒟1\mathcal{D}_{1} by specifying a normal vector field vμ=O(L)v^{\mu}=O(L) on 𝒟1\mathcal{D}_{1} such that, up to a reparameterisation of 𝒟2\mathcal{D}_{2},

X2μ(z)=X1μ(z)+vμ(z)+O(v2),\displaystyle X_{2}^{\mu}(z)=X_{1}^{\mu}(z)+v^{\mu}(z)+O(v^{2}), (3)

where zaz^{a} are defect coordinates and Xiμ(z)X_{i}^{\mu}(z) are the functions which embed the defects 𝒟i\mathcal{D}_{i} into the bulk manifold MM. Since we want to perform a systematic expansion of correlation functions in powers of vLv\sim L, it is necessary to extend (3) to higher orders in vv. In section 2.1 we provide such an extension which is both Weyl- and diffeomorphism-invariant. This is achieved by constructing a one-parameter family of deformations Xdef(z,t)X_{\text{def}}(z,t) such that, up to a reparameterisation of 𝒟2\mathcal{D}_{2},

Xdef(z,0)\displaystyle X_{\text{def}}(z,0) =X1(z),\displaystyle=X_{1}(z), (4)
Xdef(z,1)\displaystyle X_{\text{def}}(z,1) =X2(z).\displaystyle=X_{2}(z). (5)

The function Xdef(z,t)X_{\text{def}}(z,t) is defined by a Weyl-invariant partial-differential equation (19) with the initial condition

Xdefμ(z,0)\displaystyle X^{\mu}_{\text{def}}(z,0) =X1μ(z),tXdefμ(z,0)=vμ(z).\displaystyle=X^{\mu}_{1}(z),\quad\partial_{t}X^{\mu}_{\text{def}}(z,0)=v^{\mu}(z). (6)

In particular, equation (19) can be solved order-by-order in vv, providing a systematic Weyl-invariant extension of (3).444Note that we do not claim that our definition of vμv^{\mu} is unique. The non-trivial claim is that a Weyl- and diffeomorphism-invariant definition exists.

Note that this construction also gives a Weyl-covariant definition for the coupling to the displacement operator at non-linear orders. This in particular implies that there exist renormalisation schemes in which the displacement operator DD transforms as a Weyl tensor, without any anomalous terms proportional to DD.

In section 2.3 we show that the vector field vμ(z)v^{\mu}(z) on 𝒟1\mathcal{D}_{1} allows us to define a preferred Weyl frame in the neighbourhood of the two defects. It is clear that vμ(z)v^{\mu}(z) defines a length scale on 𝒟1\mathcal{D}_{1} via 2=gμνvμvν\ell^{2}=g_{\mu\nu}v^{\mu}v^{\nu} and

g^μν=2gμν\displaystyle\widehat{g}_{\mu\nu}=\ell^{-2}g_{\mu\nu} (7)

becomes a Weyl-invariant bulk metric, defined on 𝒟1\mathcal{D}_{1}. The main result of section 2.3 that the length scale \ell can be defined in a covariant way in the neighbourhood of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. This allows us to extend also g^μν\widehat{g}_{\mu\nu} to a bulk neighbourhood. This dramatically simplifies the problem of constructing SeffS_{\text{eff}} as it can now be written in terms of the Weyl-invariant metric g^μν\widehat{g}_{\mu\nu} instead of the physical metric gμνg_{\mu\nu}. We will refer to g^μν\widehat{g}_{\mu\nu} as the fusion metric.

In section 2.4 we show that the fusion metric g^μν\widehat{g}_{\mu\nu} is not completely generic and satisfies certain local constraints on 𝒟1\mathcal{D}_{1}. Up to two-derivative order the constraints are exhausted by g^μνvμvν=1\widehat{g}_{\mu\nu}v^{\mu}v^{\nu}=1, the vanishing of the trace of the second fundamental form \RomanbarII^μ=0\widehat{\Romanbar{II}}^{\mu}=0, and the vanishing of the purely normal components of the Schouten tensor. We also show that for spherical defects in flat space stronger constraints such as \RomanbarII^abμ=0\widehat{\Romanbar{II}}^{\mu}_{ab}=0 follow.

Sections 2.2 and 2.5 contain various examples of the above constructions. In particular, in section 2.2 we describe in full generality the vector field vμv^{\mu} for a pair of spherical defects (of any codimension) embedded in flat space. In section 2.5 we extend some of the examples of section 2.2 by computing the fusion metric for them.

Effective action

Using the fusion metric g^μν\widehat{g}_{\mu\nu} we can systematically construct the couplings in SeffS_{\text{eff}} in a derivative expansion suppressed by powers of LL. The leading term comes from couplings to the identity operator, and in particular from the cosmological constant term

Seffa0dpzγ^=a0dpzγpLp.\displaystyle S_{\text{eff}}\ni-a_{0}\int\,d^{p}z\,\sqrt{\widehat{\gamma}}=-a_{0}\int\,d^{p}z\,\sqrt{\gamma}\ell^{-p}\sim L^{-p}. (8)

We discuss this term in detail in section 3.4 and prove theorems 3.1 and 3.2 for the Wilson coefficient a0a_{0}. The former states that a0a_{0} is non-negative, a00a_{0}\geq 0, while the latter provides inequalities for a0a_{0} between different fusion processes.

In section 3.6 we study one-derivative identity operator couplings, which only arise for line defects in d=3d=3. We discuss their relevance in pure Chern-Simons theory. In section 3.7 we discuss the contributions of irrelevant operators. Appendix C contains a discussion of two-derivative identity couplings. For example, for line defects in d=4d=4,

Seff(2)=\displaystyle S^{(2)}_{\text{eff}}= 𝑑zγ^(a2,1^tv^tv+a2,2R^+ia2,3P^tv+a2,4Cvtvt+ia2,5C~vtvt),\displaystyle\int dz\sqrt{\widehat{\gamma}}\left(a_{2,1}\widehat{\nabla}_{t}^{\perp}v\cdot\widehat{\nabla}_{t}^{\perp}v+a_{2,2}\widehat{R}+ia_{2,3}\widehat{P}_{tv}+a_{2,4}C_{vtvt}+ia_{2,5}\widetilde{C}_{vtvt}\right), (9)

where P^μν\widehat{P}_{\mu\nu} is the Schouten tensor,555In principle, the Schouten tensor can be expressed in terms of the Ricci tensor. However, as we discuss in section 2.4, Schouten tensor generally has simpler form for fusion metrics than the Ricci tensor. CC is the Weyl tensor and tμt^{\mu} is the unit tangent to the line defect. Hats indicate that the quantities are computed in the fusion metric.

Constraints on 𝒟Σ\mathcal{D}_{\Sigma}

The position of the IR conformal defect 𝒟Σ\mathcal{D}_{\Sigma} can be taken by convention to exactly coincide with 𝒟1\mathcal{D}_{1}. This choice is possible because the effective action SeffS_{\text{eff}} will involve a coupling to the displacement operator, and we can always compensate for our choice of the position of 𝒟Σ\mathcal{D}_{\Sigma} by modifying this coupling.666An alternative choice is to set the displacement coupling to 0, in which case we would need to write down a derivative expansion for the position of 𝒟Σ\mathcal{D}_{\Sigma}. We discuss these aspects further in section 3.7.

Consider the case when 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are flat parallel defects in flat space. From the RG perspective, we treat 𝒟1𝒟2\mathcal{D}_{1}\mathcal{D}_{2} as the UV starting point of an RG flow. This configuration only preserves SO(q1)\mathrm{SO}(q-1) transverse rotations, and there is no guarantee that the symmetry will be enhanced to SO(q)\mathrm{SO}(q) at the IR fixed point. Therefore, it is possible that 𝒟Σ\mathcal{D}_{\Sigma} breaks SO(q)\mathrm{SO}(q) down to SO(q1)\mathrm{SO}(q-1). In fact, curvature couplings allow for breaking to even smaller symmetry groups. Whenever some transverse rotations are preserved at the IR fixed point, the requirement that 𝒟Σ\mathcal{D}_{\Sigma} be a stable fixed point implies that it should not have relevant operators with various transverse spins. Whenever 𝒟Σ\mathcal{D}_{\Sigma} breaks transverse rotations, it must support tilt operators and the corresponding conformal manifold, analogously to the breaking of global symmetries. We discuss these questions in detail in section 3.8.

In section 3.5 we argue that generically we expect the IR defect 𝒟Σ\mathcal{D}_{\Sigma} to be simple. We also highlight some important exceptions, including in particular the fusion of Wilson lines at weak coupling.

Cusps and Wilson lines in 𝒩=4\mathcal{N}=4 SYM

In section 4 we apply the general effective theory formalism in the context of cusp anomalous dimension. When 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} meet at an opening angle α\alpha, the resulting cusp carries a non-trivial scaling dimension Γcusp\Gamma_{\text{cusp}}. Via the exponential map, this configuration can be mapped to the cylinder ×Sd1\mathbb{R}\times S^{d-1} where the defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are inserted along \mathbb{R}, separated by angular distance α\alpha on Sd1S^{d-1}. In this picture Γcusp\Gamma_{\text{cusp}} becomes the ground state energy of the Hamiltonian HH that generates translations along \mathbb{R}. In the limit α0\alpha\to 0 the defects fuse and the effective theory can be applied. We explain how the spectrum of HH can be obtained from Rayleigh-Schrödinger perturbation theory on 𝒟Σ\mathcal{D}_{\Sigma}.

We then briefly discuss the implications of this for supersymmetric Wilson lines in planar 𝒩=4\mathcal{N}=4 SYM. We argue that the IR defect 𝒟Σ\mathcal{D}_{\Sigma} is trivial and we show that the identity operator contribution to SeffS_{\text{eff}} is responsible for the value of Γcusp\Gamma_{\text{cusp}} with error at most O(α2)O(\alpha^{2}), and possibly to all orders in α\alpha. More concretely,

Γcusp(θ,ϕ=πα)=a0(θ)α3a2,2(θ)α+O(α2),\displaystyle\Gamma_{\text{cusp}}(\theta,\phi=\pi-\alpha)=-\frac{a_{0}(\theta)}{\alpha}-3a_{2,2}(\theta)\alpha+O(\alpha^{2}), (10)

where the two-derivative Wilson coefficient a2,2a_{2,2} defined in (9). Matching the known expressions for Γcusp\Gamma_{\text{cusp}} in the ladder and in the strong-coupling limits, we compute the value of a2,2a_{2,2}.

One-point functions of bulk operators

In section 5 we apply fusion EFT to the two-point function 𝒟𝒟¯\langle\mathcal{D}\overline{\mathcal{D}}\rangle. Comparing it with the expansion of this two-point function in terms of bulk operators Gadde:2016fbj , we derive the large-Δ\Delta asymptotic of one-point functions 𝒪𝒟\langle{\mathcal{O}}\mathcal{D}\rangle of local bulk operators 𝒪{\mathcal{O}} in the presence of 𝒟\mathcal{D}. This is done for defects of any dimension pp. We also derive the spin-dependent asymptotics for line defects in d=3d=3 and d=4d=4. As an example, for line defects in d=3d=3 we find

logρ(Δ,J)8πa0Δ(112j214j6+O(j8))+O(logΔ),\displaystyle\log\rho(\Delta,J)\sim\sqrt{8\pi a_{0}\Delta}\left(1-\tfrac{1}{2}j^{2}-\tfrac{1}{4}j^{6}+O(j^{8})\right)+O(\log\Delta), (11)

where j=J/Δ1j=J/\Delta\leq 1 and ρ(Δ,J)\rho(\Delta,J) is the one-point function density for local operators (both descendants and primaries) with scaling dimension Δ\Delta and spin projection JJ. The expansion in small jj is given for illustration purposes, the full jj-dependence is determined by the transcendental equation (230).

Anomalies

In general, the conformal defects 𝒟1,𝒟2\mathcal{D}_{1},\mathcal{D}_{2} and 𝒟Σ\mathcal{D}_{\Sigma} are not exactly Weyl-invariant but can carry non-trivial Weyl anomalies. In this case, the effective action SeffS_{\text{eff}} also needs to transform non-trivially under Weyl transformations so that both sides of (2) transform in the same way. To achieve this, Weyl-anomaly matching terms need to be added to SeffS_{\text{eff}}, and we derive their general form in section 6.2. The resulting contribution is

𝒜𝒟1(g,log)+𝒜𝒟2(g,log)𝒜𝒟Σ(g,log),\displaystyle\mathcal{A}_{\mathcal{D}_{1}}(g,-\log\ell)+\mathcal{A}_{\mathcal{D}_{2}}(g,-\log\ell)-\mathcal{A}_{\mathcal{D}_{\Sigma}}(g,-\log\ell), (12)

where 𝒜(g,ω)\mathcal{A}(g,\omega) is the defect Weyl anomaly action for a finite Weyl transformation ω\omega, and \ell is evaluated on the appropriate submanifolds. As we discuss in section 7, the existence of the scale \ell is crucial for this construction, and the analogous approach doesn’t work in the case of local operator OPE.

Generally speaking, the Weyl anomaly 𝒜𝒟Σ\mathcal{A}_{\mathcal{D}_{\Sigma}} belongs to a class that is slightly more general than usually considered in the literature (see Chalabi:2021jud for a review) since the defect 𝒟Σ\mathcal{D}_{\Sigma} might break some transverse rotations. In sections 6.3 and 6.4 we begin the classification of Weyl anomalies for conformal defects that break the transverse rotations down to SO(q1)\mathrm{SO}(q-1). We find that such line defects admit a new B-type Weyl anomaly in d=3d=3 and no Weyl anomalies in d>3d>3. For surface defects, we find 5 new B-type terms in d=4d=4.

2 Conformal geometry of a pair of defects

In subsection 2.1 we derive a way of parameterizing the relative position of two conformal defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} in a Weyl-invariant way. In subsection 2.3 we show that similar ideas lead to a definition of a canonical Weyl frame in the neighbourhood of 𝒟1,𝒟2\mathcal{D}_{1},\mathcal{D}_{2}. In section 2.4 we discuss the special relations satisfied in this canonical Weyl frame. We give explicit examples of these constructions in subsections 2.2 and 2.5.

The defects we consider have dimension p>0p>0 and co-dimension qq, with p+q=dp+q=d the dimension of the bulk. We assume that d>2d>2. We work in Euclidean signature and on a bulk manifold MM with a general metric gg. We assume that the defect manifold is NN for both 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}, embedded into MM via X1:NMX_{1}:N\to M and X2:NMX_{2}:N\to M respectively. We use Greek letters for bulk indices (xμx^{\mu}) and Latin letters for defect indices (zaz^{a}). We will use xx to denote points in MM and zz to denote points in NN, and we will often use 𝒟i\mathcal{D}_{i} to refer to the images Xi(N)X_{i}(N). To declutter the notation, and when there is no fear of confusion, we may use zz to refer to the point Xi(z)X_{i}(z) on 𝒟i\mathcal{D}_{i}. Some additional comments regarding the geometry are given in appendix A.

2.1 A Weyl-invariant displacement

We focus on describing the position of 𝒟2\mathcal{D}_{2} relative to 𝒟1\mathcal{D}_{1} when the defects are close to one another. We will do this by describing the possible deformations of the embedding function X1X_{1} and then determining the deformation that yields X2X_{2} (modulo defect diffeomorphisms).

To the leading order, a deformation of X1X_{1} can be parameterised by a bulk vector field vμ(z)v^{\mu}(z) defined on 𝒟1\mathcal{D}_{1},777Formally, one would describe vv as a section of the vector bundle over NN that one obtains from the tangent bundle TMTM by pulling back along X1X_{1}. To simplify the discussion, we will appeal to more informal descriptions of various quantities throughout the paper. so that the deformed embedding function X1,vX_{1,v} is defined as

X1,vμ(z)=X1μ(z)+vμ(z)+O(v2).\displaystyle X_{1,v}^{\mu}(z)=X_{1}^{\mu}(z)+v^{\mu}(z)+O(v^{2}). (13)

Components of vμ(z)v^{\mu}(z) tangent to 𝒟1=X1(N)\mathcal{D}_{1}=X_{1}(N) do not affect the deformed defect X1,v(N)X_{1,v}(N) at this order because they can be absorbed into diffeomorphisms of NN. We can fix this redundancy by requiring that vμv^{\mu} be normal to 𝒟1\mathcal{D}_{1}.

This parameterisation is explicitly Weyl-invariant. Indeed, Weyl transformations only affect the metric which enters here only through the condition that vv is orthogonal to 𝒟1\mathcal{D}_{1}, and angles are unchanged by Weyl rescalings. However, so far we have only defined the deformation X1,vX_{1,v} to the leading order in vv. Our goal, on the other hand, is to define the deformation to all non-linear orders in the deformation parameter vv.

A standard way to extend the definition (13) to non-linear orders is by using affine geodesics with initial velocity vμv^{\mu}. For this, we first introduce a tt-dependent deformation Xdef(z,t)X_{\text{def}}(z,t), so that the final deformation is given by

X1,vμ(z)Xdef(z,t=1).\displaystyle X_{1,v}^{\mu}(z)\equiv X_{\text{def}}(z,t=1). (14)

We could then require that for a fixed zz, Xdef(z,t)X_{\text{def}}(z,t) satisfies the geodesic equation

t2Xdefμ=t2Xdefμ+ΓαβμtXdefαtXdefβ=0,(not our definition)\displaystyle\nabla_{t}^{2}X^{\mu}_{\text{def}}=\partial_{t}^{2}X_{\text{def}}^{\mu}+\Gamma^{\mu}_{\alpha\beta}\partial_{t}X^{\alpha}_{\text{def}}\partial_{t}X^{\beta}_{\text{def}}=0,\quad{\color[rgb]{.5,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,.5,.5}\pgfsys@color@gray@stroke{.5}\pgfsys@color@gray@fill{.5}\text{(not our definition)}} (15)

subject to the initial conditions

Xdefμ(z,0)=X1μ(z),tXdefμ(z,0)=vμ(z).\displaystyle X^{\mu}_{\text{def}}(z,0)=X^{\mu}_{1}(z),\quad\partial_{t}X^{\mu}_{\text{def}}(z,0)=v^{\mu}(z). (16)

This is essentially a version of Riemann normal coordinates in a neighbourhood of a submanifold. While this definition is diff-invariant,888Something like X1,vμ(z)X1μ(z)+vμ(z)X_{1,v}^{\mu}(z)\equiv X_{1}^{\mu}(z)+v^{\mu}(z) is, on the other hand, not even diff-invariant. it fails to be Weyl-invariant since the Christoffel symbols Γ\Gamma transform inhomogeneously as

δωΓρμν\displaystyle\delta_{\omega}\Gamma^{\rho}{}_{\mu\nu} =δμρνω+δνρμωgμνρω.\displaystyle=\delta^{\rho}_{\mu}\partial_{\nu}\omega+\delta^{\rho}_{\nu}\partial_{\mu}\omega-g_{\mu\nu}\partial^{\rho}\omega. (17)

It is easy to check that the problem arises already in the O(v2)O(v^{2}) term in X1,vμX_{1,v}^{\mu} resulting from this definition,

X1,vμ=X1μ+vμ12Γαβμvαvβ+O(v3).\displaystyle X^{\mu}_{1,v}=X_{1}^{\mu}+v^{\mu}-\frac{1}{2}\Gamma^{\mu}_{\alpha\beta}v^{\alpha}v^{\beta}+O(v^{3}). (18)

To overcome this problem we use a Weyl-invariant equation in place of (15),

t2Xdefμ+2ptXdefμtXdefσ\RomanbarIIσ1p(tXdef)2\RomanbarIIμ+12aXdefμa(tXdef)2=0,\displaystyle\nabla_{t}^{2}X^{\mu}_{\text{def}}+\frac{2}{p}\partial_{t}X^{\mu}_{\text{def}}\partial_{t}X^{\sigma}_{\text{def}}\Romanbar{II}_{\sigma}-\frac{1}{p}(\partial_{t}X_{\text{def}})^{2}\Romanbar{II}^{\mu}+\frac{1}{2}\partial_{a}X^{\mu}_{\text{def}}\partial^{a}(\partial_{t}X_{\text{def}})^{2}=0, (19)

which is now a partial-differential equation in z,tz,t.999The intuition behind this equation is as follows. The equation t2Xdef=0\nabla_{t}^{2}X_{\text{def}}=0 transforms by first derivatives of the Weyl factor on the defect Xdef(,t)X_{\text{def}}(\cdot,t). So, if we can fix the first derivatives of a Weyl scale on this defect, we can turn t2Xdef=0\nabla_{t}^{2}X_{\text{def}}=0 into a conformally-invariant equation. We first restrict to the so-called minimal scales in which \RomanbarIIμ=0\Romanbar{II}^{\mu}=0, which fixes the normal derivatives of the scale. We then fix the tangential derivatives of the scale by requiring that (tXdef)2=1(\partial_{t}X_{\text{def}})^{2}=1. It is then easy to see that in this scale (19) reduces to t2Xdef=0\nabla_{t}^{2}X_{\text{def}}=0. Here, \RomanbarIIμ=\RomanbarIIabμg^ab\Romanbar{II}^{\mu}=\Romanbar{II}^{\mu}_{ab}\widehat{g}^{ab} is the trace of the second fundamental form calculated for the embedding Xdef(,t)X_{\text{def}}(\cdot,t). This equation is to be solved with initial conditions (16). Using the transformation rules given in appendix A, it is easy to check that the equation is invariant under Weyl transformations as long as long as

tXdefaXdef=0.\displaystyle\partial_{t}X_{\text{def}}\cdot\partial_{a}X_{\text{def}}=0. (20)

This condition is in turn true because it is satisfied by the initial condition (16) and equation (19) implies that

t(tXdefaXdef)=2ptXdefν\RomanbarIIν(tXdefaXdef).\displaystyle\partial_{t}\left(\partial_{t}X_{\text{def}}\cdot\partial_{a}X_{\text{def}}\right)=-\frac{2}{p}\partial_{t}X_{\text{def}}^{\nu}\Romanbar{II}_{\nu}\left(\partial_{t}X_{\text{def}}\cdot\partial_{a}X_{\text{def}}\right). (21)

The solution to (19) can be systematically constructed order by order in vv. For example, the first two orders yield

X1,vμ=\displaystyle X^{\mu}_{1,v}= X1μ+vμ12Γαβμvαvβ1pvμvσ\RomanbarIIσ+12pv2\RomanbarIIμ14aX1μa(v2)+O(v3).\displaystyle X_{1}^{\mu}+v^{\mu}-\frac{1}{2}\Gamma^{\mu}_{\alpha\beta}v^{\alpha}v^{\beta}-\frac{1}{p}v^{\mu}v^{\sigma}\Romanbar{II}_{\sigma}+\frac{1}{2p}v^{2}\Romanbar{II}^{\mu}-\frac{1}{4}\partial_{a}X^{\mu}_{1}\partial^{a}(v^{2})+O(v^{3}). (22)

Higher orders can be easily obtained by taking further tt-derivatives of (19) at t=0t=0. Unlike (18), this expansion is Weyl-invariant. This has two important consequences.

Firstly, this construction allows us to parameterise the relative position of the defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} using a vector field vμv^{\mu}. For this, we search for vμv^{\mu} such that

X1,v(z)=X2(φ(z))\displaystyle X_{1,v}(z)=X_{2}(\varphi(z)) (23)

for some diffeomorphism φ:NN\varphi:N\to N. Note that by construction, vμv^{\mu} does not transform under Weyl transformations. On the other hand, its length

(z)gμνvμvν\displaystyle\ell(z)\equiv\sqrt{g_{\mu\nu}v^{\mu}v^{\nu}} (24)

transforms with Weyl weight 11, as expected from an infinitesimal distance. The function (z)\ell(z) formalises the idea of “coordinate-dependent” distance between the two defects and will feature prominently (alongside vμv^{\mu}) in the effective actions that we study in section 3.

Secondly, this proves the existence of Weyl-covariant schemes for the coupling of the displacement operator. Normally, the displacement operator DμD_{\mu} is defined so that

𝒟1[edpzγvμDμ],\displaystyle\langle\mathcal{D}_{1}[e^{-\int d^{p}z\sqrt{\gamma}v^{\mu}D_{\mu}}]\cdots\rangle, (25)

where γ\gamma is the defect metric, computes a correlator with 𝒟1\mathcal{D}_{1} displaced to X1=X1+v+O(v2)X_{1}^{\prime}=X_{1}+v+O(v^{2}) (see, for example Billo:2016cpy ). Specifying this displacement to higher orders in vv is equivalent to defining a renormalisation scheme for correlation functions of DμD_{\mu}. The fact that we are able to make the Weyl-invariant definition X1X1,vX_{1}^{\prime}\equiv X_{1,v} implies101010For this it is important that X1,vX_{1,v} has a local series expansion in terms of vv and its derivatives, see (22). If this were violated, we wouldn’t expect to be able to use vv as the expansion parameter in conformal perturbation theory. that the corresponding renormalisation scheme contains no scale anomalies in contact terms of DμD_{\mu} with itself. Put differently, the action dpzg^vμDμ\int d^{p}z\sqrt{\widehat{g}}v^{\mu}D_{\mu} can be made Weyl-invariant at quantum level.

We should note that other Weyl-invariant definitions of X1,vX_{1,v} might be possible. We will not rely on any special properties of the above definition beyond those already mentioned.

2.2 Example: spherical defects in flat space I

The discussion in section 2.1 applies in general curved backgrounds and for arbitrarily shaped defects. If the spacetime manifold is flat M=dM=\mathbb{R}^{d} (more accurately, its conformal compactification SdS^{d}) and the defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are flat or spherical, the definitions can be made much more explicit, which is the goal of this section. In what follows we set gμν=δμνg_{\mu\nu}=\delta_{\mu\nu}, and we use “spherical” to refer to both spherical and flat defects (the latter understood as spheres of infinite radius) of any dimension.

To begin, we first fix our conventions for conformal Killing vectors (CKVs). Let ξμ(x)\xi^{\mu}(x) be a CKV on M=dM=\mathbb{R}^{d}, so that it satisfies the equation

μξν+νξμ=2dδμν(ξ).\displaystyle\partial_{\mu}\xi_{\nu}+\partial_{\nu}\xi_{\mu}=\frac{2}{d}\delta_{\mu\nu}(\partial\cdot\xi). (26)

We define the following basis of Killing vectors

𝐝μ(x)\displaystyle\mathbf{d}^{\mu}(x) =xμ,\displaystyle=x^{\mu}, (27)
(𝐩ν)μ(x)\displaystyle(\mathbf{p}_{\nu})^{\mu}(x) =δνμ,\displaystyle=\delta^{\mu}_{\nu}, (28)
(𝐤ν)μ(x)\displaystyle(\mathbf{k}_{\nu})^{\mu}(x) =2xμxνδνμx2,\displaystyle=2x^{\mu}x_{\nu}-\delta^{\mu}_{\nu}x^{2}, (29)
(𝐦ρσ)μ(x)\displaystyle(\mathbf{m}_{\rho\sigma})^{\mu}(x) =δρμxσδσμxρ.\displaystyle=\delta^{\mu}_{\rho}x_{\sigma}-\delta^{\mu}_{\sigma}x_{\rho}. (30)

The general solution to (26) is a linear combination of these basis elements,

ξ=λ𝐝+aμ𝐩μ+bμ𝐤μ+12wμν𝐦μν.\displaystyle\xi=\lambda\mathbf{d}+a^{\mu}\mathbf{p}_{\mu}+b^{\mu}\mathbf{k}_{\mu}+\frac{1}{2}w^{\mu\nu}\mathbf{m}_{\mu\nu}. (31)

The CKVs form a Lie algebra under the Lie bracket of vector fields – the Lie algebra of conformal transformations of d\mathbb{R}^{d}, and the non-zero values of the corresponding Killing form are

𝐝,𝐝=2d,𝐩μ,𝐤ν=4dδμν,𝐦μν,𝐦ρσ=2d(δμρδνσδμσδνρ).\displaystyle\langle\mathbf{d},\mathbf{d}\rangle=2d,\quad\langle\mathbf{p}_{\mu},\mathbf{k}_{\nu}\rangle=-4d\delta_{\mu\nu},\quad\langle\mathbf{m}_{\mu\nu},\mathbf{m}_{\rho\sigma}\rangle=-2d\left(\delta_{\mu\rho}\delta_{\nu\sigma}-\delta_{\mu\sigma}\delta_{\nu\rho}\right). (32)

Recall that φ=eξ\varphi=e^{\xi} is a conformal transformation that is constructed as follows. To compute φ(x0)\varphi(x_{0}) we solve the differential equation

x˙μ(t)=ξμ(x(t))\displaystyle\dot{x}^{\mu}(t)=\xi^{\mu}(x(t)) (33)

with the initial condition x(0)=x0x(0)=x_{0} and then set φ(x0)=x(1)\varphi(x_{0})=x(1).

We now fix a CKV ξ\xi and define

Xdef(z,t)=etξ(X1(z)),\displaystyle X_{\text{def}}(z,t)=e^{t\xi}(X_{1}(z)), (34)

and suppose that Xdef(z,t)X_{\text{def}}(z,t) solves (19) at t=0t=0. We will discuss this condition later. We further assume that ξμ\xi^{\mu}, when restricted to 𝒟1\mathcal{D}_{1}, is orthogonal to 𝒟1\mathcal{D}_{1} so that v=ξ|𝒟1v=\xi|_{\mathcal{D}_{1}} provides an initial condition (16) of the kind discussed in section 2.1. We will see later that there is a healthy supply of such ξ\xi for spherical 𝒟1\mathcal{D}_{1}. We claim that (34) then solves (19) for all times tt.

To prove our claim, we first show that ξμ\xi^{\mu} is orthogonal to the image of Xdef(,t)X_{\text{def}}(\cdot,t) for any tt. Indeed, notice that etξe^{t\xi} is a conformal transformation, and therefore ξ\xi being orthogonal to etξ(X1())e^{t\xi}(X_{1}(\cdot)) is equivalent to d(etξ)(ξ)d(e^{-t\xi})(\xi) being orthogonal to X1()X_{1}(\cdot). Here, d(etξ)(ξ)d(e^{-t\xi})(\xi) is the pushforward of ξ\xi by the diffeomorphism etξe^{-t\xi} and can be written as etξξe^{-t\mathcal{L}_{\xi}}\xi, where ξ\mathcal{L}_{\xi} is the Lie derivative. Since ξξ=[ξ,ξ]=0\mathcal{L}_{\xi}\xi=[\xi,\xi]=0, it follows that d(etξ)(ξ)=ξd(e^{-t\xi})(\xi)=\xi and therefore d(etξ)(ξ)d(e^{-t\xi})(\xi) is indeed orthogonal to X1()X_{1}(\cdot) by our assumptions.

Equations (33) and (34) show that ξ(Xdef(z,t))\xi(X_{\text{def}}(z,t)) coincides with tX1(z,t)\partial_{t}X_{1}(z,t). Therefore, the above orthogonality statement implies, following the discussion in section 2.1, that the left-hand side of (19) is conformally-invariant when evaluated on Xdef(z,t)X_{\text{def}}(z,t). By the definition (34), time evolution of Xdef(z,t)X_{\text{def}}(z,t) is precisely the conformal transformation etξe^{t\xi}. Therefore, similarly to the above argument for orthogonality, the fact that (19) is satisfied for t=0t=0 implies that it is satisfied for all tt.

Let us consider a concrete example where 𝒟1\mathcal{D}_{1} is a flat defect passing through x=0x=0, and spanning the coordinate directions 1,,p1,\cdots,p. In this case, (19) becomes at t=0t=0

t2Xdefμ+12aXdefμa(tXdef)2=0.\displaystyle\partial^{2}_{t}X_{\text{def}}^{\mu}+\frac{1}{2}\partial_{a}X_{\text{def}}^{\mu}\partial^{a}(\partial_{t}X_{\text{def}})^{2}=0. (35)

In terms of ξ\xi and X1X_{1} this becomes

ξννξμ+12aX1μaξ2=0.\displaystyle\xi^{\nu}\partial_{\nu}\xi^{\mu}+\frac{1}{2}\partial_{a}X_{1}^{\mu}\partial^{a}\xi^{2}=0. (36)

The condition that ξ\xi is orthogonal to 𝒟1\mathcal{D}_{1} allows for the following choices for ξ\xi:

  • translations 𝐩μ\mathbf{p}_{\mu} normal to 𝒟1\mathcal{D}_{1}, i.e. μ=p+1,,d\mu=p+1,\cdots,d,

  • special conformal transformations 𝐤μ\mathbf{k}_{\mu} normal to 𝒟1\mathcal{D}_{1}, i.e. μ=p+1,,d\mu=p+1,\cdots,d,

  • rotations 𝐦μν\mathbf{m}_{\mu\nu} with at least one index normal to the defect, i.e. μ=p+1,,d\mu=p+1,\cdots,d or ν=p+1,,d\nu=p+1,\cdots,d, or both.

By an explicit calculation, one can check that if the rotations 𝐦μν\mathbf{m}_{\mu\nu} with both indices normal to the defect are excluded, then any linear combination of the remaining CKVs satisfies (36). We will refer to such ξ\xi as standard deformations of 𝒟1\mathcal{D}_{1}. Therefore, the standard deformations are spanned by 𝐩μ,𝐤μ,𝐦μν\mathbf{p}_{\mu},\mathbf{k}_{\mu},\mathbf{m}_{\mu\nu} with μ=p+1,,d\mu=p+1,\cdots,d and ν=1,,p\nu=1,\cdots,p.

Let 𝔥𝔰𝔬(d+1,1)\mathfrak{h}\subseteq\mathfrak{so}(d+1,1) be the subalgebra of conformal transformations which leave 𝒟1\mathcal{D}_{1} invariant and HH be the corresponding subgroup of SO(d+1,1)\mathrm{SO}(d+1,1). That is,

𝔥𝔰𝔬(p+1,1)𝔰𝔬(dp)\displaystyle\mathfrak{h}\simeq\mathfrak{so}(p+1,1)\oplus\mathfrak{so}(d-p) (37)

and 𝔥\mathfrak{h} is spanned by 𝐝,𝐩μ,𝐤μ,𝐦μν\mathbf{d},\mathbf{p}_{\mu},\mathbf{k}_{\mu},\mathbf{m}_{\mu\nu} with μ,ν1,,p\mu,\nu\in 1,\cdots,p and by 𝐦μν\mathbf{m}_{\mu\nu} with μ,ν=p+1,,d\mu,\nu=p+1,\cdots,d. It is then easy to check that the standard deformations form precisely the subspace 𝔥\mathfrak{h}^{\perp} orthogonal to 𝔥\mathfrak{h} with respect to the Killing form (32). This subspace can be identified with the quotient 𝔰𝔬(d+1,1)/𝔥\mathfrak{so}(d+1,1)/\mathfrak{h}. In other words, at least infinitesimally, the standard deformations generate SO(d+1,1)/H\mathrm{SO}(d+1,1)/H which is the space of all spherical defects of dimension pp.

This implies that the solutions (34) constructed above describe all possible deformations of 𝒟1\mathcal{D}_{1} which preserve the property of it being spherical. In other words, the position of a spherical defect 𝒟2\mathcal{D}_{2} relative to a spherical defect 𝒟1\mathcal{D}_{1} can always be described by v=ξ|𝒟1v=\xi|_{\mathcal{D}_{1}} where ξ\xi is a standard deformation of 𝒟1\mathcal{D}_{1}. Furthermore, using the invariance properties of the Killing form on 𝔰𝔬(d+1,1)\mathfrak{so}(d+1,1) we can see that the space of standard deformations of a spherical defect 𝒟1\mathcal{D}_{1} is always111111I.e. not just in the above example of a flat defect. 𝔥\mathfrak{h}^{\perp}, where 𝔥\mathfrak{h} is the subalgebra that stabilises 𝒟1\mathcal{D}_{1}.

Refer to caption
Figure 2: An example configuration of line defects in d=3d=3. The thick blue circle is 𝒟1\mathcal{D}_{1}, the thick red circle is 𝒟2\mathcal{D}_{2}. The thin lines in between represent different values of tt in Xdef(z,t)X_{\text{def}}(z,t) defined in section 2.1, ranging from 0 to 11 with step 0.10.1.

Example: line defects in d=3d=3

A pair of line defects in d=3d=3 can always be mapped to the configuration where 𝒟1\mathcal{D}_{1} is a circle of radius 11 in the plane 1-2 with its centre at 0, while 𝒟2\mathcal{D}_{2} is smaller circle obtained by applying a dilatation and a rotation in the plane 131-3, see figure 2. Specifically, 𝒟2\mathcal{D}_{2} is obtained from 𝒟1\mathcal{D}_{1} by action with (note that [𝐝,𝐦13]=0[\mathbf{d},\mathbf{m}_{13}]=0)

elogr𝐝θ𝐦13=elogr𝐝eθ𝐦13,\displaystyle e^{\log r\mathbf{d}-\theta\mathbf{m}_{13}}=e^{\log r\mathbf{d}}e^{-\theta\mathbf{m}_{13}}, (38)

where r<1r<1 is the radius of 𝒟2\mathcal{D}_{2} and θ\theta is an angle. The intersections of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} with the x2=0x^{2}=0 plane mark four points which kinematically behave similarly to operator insertions in a four-point function.121212The plane x2=0x^{2}=0 is the unique plane or sphere which meets both defects at right angles.

If we introduce a complex coordinate x1+ix3x^{1}+ix^{3} in this plane, then 𝒟1\mathcal{D}_{1} intersects it at points ±1\pm 1, while 𝒟2\mathcal{D}_{2} intersects it at ±ρ\pm\rho, where

ρ=reiθ.\displaystyle\rho=re^{i\theta}. (39)

In these kinematics ρ\rho is a cross-ratio, and it is completely analogous to the radial ρ\rho cross-ratio in four-point functions Hogervorst:2013sma .

The subalgebra 𝔥\mathfrak{h} that preserves 𝒟1\mathcal{D}_{1} is given by

𝔥=Span{𝐦13,𝐩1𝐤1,𝐩2𝐤2,𝐩3+𝐤3}.\displaystyle\mathfrak{h}=\mathrm{Span}\{\mathbf{m}_{13},\mathbf{p}_{1}-\mathbf{k}_{1},\mathbf{p}_{2}-\mathbf{k}_{2},\mathbf{p}_{3}+\mathbf{k}_{3}\}. (40)

We can easily see from (32) that both 𝐝\mathbf{d} and 𝐦13\mathbf{m}_{13} belong to 𝔥\mathfrak{h}^{\perp}. Therefore, the CKV

ξ=logr𝐝θ𝐦13\displaystyle\xi=\log r\mathbf{d}-\theta\mathbf{m}_{13} (41)

is a standard deformation and we already know that eξe^{\xi} maps 𝒟1\mathcal{D}_{1} to 𝒟2\mathcal{D}_{2}.

Restricting ξ\xi to 𝒟1\mathcal{D}_{1} we find

v=ξ|𝒟1=logre^r+θcosϕe^3,\displaystyle v=\xi|_{\mathcal{D}_{1}}=\log r\widehat{e}_{r}+\theta\cos\phi\widehat{e}_{3}, (42)

where e^r\widehat{e}_{r} and e^3\widehat{e}_{3} are unit vectors in the radial and the second directions, respectively, while ϕ\phi is an angular coordinate on 𝒟1\mathcal{D}_{1} such that x1=cosϕx_{1}=\cos\phi and x2=sinϕx_{2}=\sin\phi. The function \ell takes on the defect the values

(ϕ)=(logr)2+θ2cos2ϕ.\displaystyle\ell(\phi)=\sqrt{(\log r)^{2}+\theta^{2}\cos^{2}\phi}. (43)

Example: dimension-pp defects

The previous example can be generalised to the situation when 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are pp-dimensional and the bulk is dd-dimensional. Specifically, we take 𝒟1\mathcal{D}_{1} to be a pp-dimensional spherical defect of radius 11 centred at 0 and lying in the subspace spanned by the coordinate directions 1,,p+11,\cdots,p+1. The defect 𝒟2\mathcal{D}_{2} is then obtained by the action of

elogr𝐝iθi𝐦i,p+i+1\displaystyle e^{\log r\mathbf{d}-\sum_{i}\theta_{i}\mathbf{m}_{i,p+i+1}} (44)

where the sum is over i=1,,m=min{p+1,dp1}i=1,\cdots,m=\min\{p+1,d-p-1\}. Note that by construction, all the generators in the exponential commute with each other. Furthermore, the total number of parameters r,θ1,,θmr,\theta_{1},\cdots,\theta_{m} is m+1=min{d+2q,q}m+1=\min\{d+2-q,q\}. This is the same as the number of cross-ratios for a pair of pp-dimensional defects Gadde:2016fbj .

Generalizing the previous example, we have 𝔥=𝔰𝔬(1,p+1)𝔰𝔬(q)\mathfrak{h}=\mathfrak{so}(1,p+1)\oplus\mathfrak{so}(q). The subalgebra 𝔰𝔬(p+1)𝔰𝔬(1,p+1)\mathfrak{so}(p+1)\subseteq\mathfrak{so}(1,p+1) is given by rotations in the first p+1p+1 directions, while the remaining p+1p+1 generators of 𝔰𝔬(1,p+1)\mathfrak{so}(1,p+1) are given by 𝐩μ𝐤μ\mathbf{p}_{\mu}-\mathbf{k}_{\mu} with μ{1,,p+1}\mu\in\{1,\cdots,p+1\}. In other words,

𝔰𝔬(1,p+1)=Span{𝐩μ𝐤μ|μ{1,,p+1}}+Span{𝐦μν|μ,ν{1,,p+1}}.\displaystyle\mathfrak{so}(1,p+1)=\mathrm{Span}\{\mathbf{p}_{\mu}-\mathbf{k}_{\mu}|\mu\in\{1,\cdots,p+1\}\}+\mathrm{Span}\{\mathbf{m}_{\mu\nu}|\mu,\nu\in\{1,\cdots,p+1\}\}. (45)

Similarly 𝔰𝔬(q1)𝔰𝔬(q)\mathfrak{so}(q-1)\subset\mathfrak{so}(q) is given by rotations in the last q1q-1 directions, while the remaining q1q-1 generators of 𝔰𝔬(q)\mathfrak{so}(q) are given by 𝐩μ+𝐤μ\mathbf{p}_{\mu}+\mathbf{k}_{\mu} with μ{p+2,,d}\mu\in\{p+2,\cdots,d\}, i.e.

𝔰𝔬(q)=Span{𝐩μ+𝐤μ|μ{p+2,,d}}+Span{𝐦μν|μ,ν{p+2,,d}}.\displaystyle\mathfrak{so}(q)=\mathrm{Span}\{\mathbf{p}_{\mu}+\mathbf{k}_{\mu}|\mu\in\{p+2,\cdots,d\}\}+\mathrm{Span}\{\mathbf{m}_{\mu\nu}|\mu,\nu\in\{p+2,\cdots,d\}\}. (46)

This leaves

𝔥=\displaystyle\mathfrak{h}^{\perp}= Span{𝐝}+Span{𝐩μ+𝐤μ|μ{1,,p+1}}+Span{𝐩μ𝐤μ|μ{p+2,,d}}\displaystyle\;\mathrm{Span}\{\mathbf{d}\}+\mathrm{Span}\{\mathbf{p}_{\mu}+\mathbf{k}_{\mu}|\mu\in\{1,\cdots,p+1\}\}+\mathrm{Span}\{\mathbf{p}_{\mu}-\mathbf{k}_{\mu}|\mu\in\{p+2,\cdots,d\}\}
+Span{𝐦μν|μ{1,,p+1},ν{p+2,,d}}.\displaystyle+\mathrm{Span}\{\mathbf{m}_{\mu\nu}|\mu\in\{1,\cdots,p+1\},\nu\in\{p+2,\cdots,d\}\}. (47)

In particular, the CKV appearing in (44) is in 𝔥\mathfrak{h}^{\perp}, and is therefore a standard deformation. So, we can take

ξ=logr𝐝iθi𝐦i,p+i+1,\displaystyle\xi=\log r\mathbf{d}-\sum_{i}\theta_{i}\mathbf{m}_{i,p+i+1}, (48)

and therefore

v=ξ|𝒟1=logre^r+iθixie^p+i+1.\displaystyle v=\xi|_{\mathcal{D}_{1}}=\log r\widehat{e}_{r}+\sum_{i}\theta_{i}x^{i}\widehat{e}_{p+i+1}. (49)

Here, xix^{i} are the bulk coordinates of a point on 𝒟1\mathcal{D}_{1}. The function \ell is then given by (on 𝒟1\mathcal{D}_{1})

(z)=(logr)2+iθi2(xi)2.\displaystyle\ell(z)=\sqrt{(\log r)^{2}+\textstyle\sum_{i}\theta_{i}^{2}(x^{i})^{2}}. (50)
Refer to caption
Figure 3: An example configuration of line defects in d=3d=3. The thick blue circle is 𝒟1\mathcal{D}_{1}, the thick red circle is 𝒟2\mathcal{D}_{2}. The thin lines in between represent different values of tt in Xdef(z,t)X_{\text{def}}(z,t) defined in section 2.1, ranging from 0 to 11 with step 0.10.1. In this picture, L=4L=4 is not small to show the path of the deformation more clearly.

Example: parallel dimension-pp spherical defects

We now consider the case when 𝒟1\mathcal{D}_{1} is as above, but 𝒟2\mathcal{D}_{2} is obtained from it by a translation by LL in the coordinate xdx^{d}. This is conformally equivalent to a configuration in the previous example with θi=0\theta_{i}=0 and

r=exp(2arcsinhL2)=(1+L24+L2)2.\displaystyle r=\exp\left(-2\,\mathrm{arcsinh}\frac{L}{2}\right)=\left(\sqrt{1+\tfrac{L^{2}}{4}}+\tfrac{L}{2}\right)^{-2}. (51)

However, it is instructive to consider this configuration directly.

Note that while 𝒟2\mathcal{D}_{2} can be obtained by applying eL𝐩de^{L\mathbf{p}_{d}} to 𝒟1\mathcal{D}_{1}, L𝐩dL\mathbf{p}_{d} is not a standard deformation. This follows from the characterisation of 𝔥\mathfrak{h}^{\perp} in (2.2). Instead, we have to take

ξ=αtanhα2𝐝+α2coshα2(𝐩d𝐤d)\displaystyle\xi=\alpha\tanh\tfrac{\alpha}{2}\mathbf{d}+\frac{\alpha}{2\cosh\frac{\alpha}{2}}(\mathbf{p}_{d}-\mathbf{k}_{d}) (52)

where α\alpha is such that

L=2sinhα2.\displaystyle L=2\sinh\tfrac{\alpha}{2}. (53)

That this works can be verified by an explicit calculation using embedding space methods. One can also check that as tt goes from 0 to 11, the transformation etξe^{t\xi} moves 𝒟1\mathcal{D}_{1} into 𝒟2\mathcal{D}_{2} along the unique (p+1)(p+1)-dimensional sphere on which both 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} lie, see figure 3.

This choice of ξ\xi gives

v=ξ|𝒟1=αtanhα2e^r+αcoshα2e^d\displaystyle v=\xi|_{\mathcal{D}_{1}}=\alpha\tanh\tfrac{\alpha}{2}\widehat{e}_{r}+\frac{\alpha}{\cosh\frac{\alpha}{2}}\widehat{e}_{d} (54)

and

(z)=α.\displaystyle\ell(z)=\alpha. (55)

Example: parallel dimension-pp flat defects

Next, we consider the case when 𝒟1\mathcal{D}_{1} is a flat defect passing through x=0x=0, and spanning the coordinate directions 1,,p1,\cdots,p, but 𝒟2\mathcal{D}_{2} is again obtained from it by a translation by LL in the coordinate xdx^{d}.

We represent this translation by setting ξ=L𝐩d\xi=L\mathbf{p}_{d}, which we already know is a standard deformation. This then clearly gives that

vμ=Le^d\displaystyle v^{\mu}=L\,\widehat{e}_{d} (56)

and

(z)=L.\displaystyle\ell(z)=L. (57)

Example: parallel flat line defects in ×Sd1\mathbb{R}\times S^{d-1}

Finally, we consider the case when 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are flat parallel line defects (so p=1p=1) in ×Sd1\mathbb{R}\times S^{d-1}, placed along the time direction \mathbb{R}. This setup is conformally equivalent to a setup in d{0}\mathbb{R}^{d}\setminus\{0\} via the exponential map exp:×Sd1d{0}\exp:\mathbb{R}\times S^{d-1}\to\mathbb{R}^{d}\setminus\{0\}, defined as

exp(τ,Ω)eτΩ\displaystyle\exp(\tau,\Omega)\equiv e^{\tau}\Omega (58)

for τ\tau\in\mathbb{R} and ΩSd1\Omega\in{S}^{d-1} (parameterised as a unit vector in d\mathbb{R}^{d}). In d{0}\mathbb{R}^{d}\setminus\{0\}, the defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} become rays starting at 0, and pointing in two different directions Ω1\Omega_{1} and Ω2\Omega_{2}, respectively.

Without loss of generality, we can assume that

Ω1=(1,0,0,,0),Ω2=(cosα,sinα,0,0,0),\displaystyle\Omega_{1}=(1,0,0,\dots,0),\qquad\Omega_{2}=(\cos\alpha,\sin\alpha,0,0,\dots 0), (59)

where α\alpha is the angle between Ω1\Omega_{1} and Ω2\Omega_{2}. From the preceding discussion, we know that in the d{0}\mathbb{R}^{d}\setminus\{0\} picture, ξ=α𝐦21\xi=\alpha\mathbf{m}_{21} is a standard deformation for 𝒟1\mathcal{D}_{1}. On the other hand, eξe^{\xi} is a rotation by angle α\alpha which maps Ω1\Omega_{1} to Ω2\Omega_{2}, and therefore 𝒟1\mathcal{D}_{1} to 𝒟2\mathcal{D}_{2}. Thus, in d{0}\mathbb{R}^{d}\setminus\{0\},

vμ=ξμ|𝒟1=αx1δ2μ.\displaystyle v^{\mu}=\xi^{\mu}|_{\mathcal{D}_{1}}=\alpha x^{1}\delta^{\mu}_{2}. (60)

Under exp1\exp^{-1}, this maps to a time-independent vector of length α\alpha, which lies along the spatial directions on ×Sd1\mathbb{R}\times S^{d-1} and points along the great circle from 𝒟1\mathcal{D}_{1} to 𝒟2\mathcal{D}_{2}. Therefore, in the Weyl frame of ×Sd1\mathbb{R}\times S^{d-1},

(z)=α.\displaystyle\ell(z)=\alpha. (61)

2.3 Fusion metric

In the previous subsection we have defined a vector field vμ(z)v^{\mu}(z) and the corresponding length scale (z)=v2(z)\ell(z)=\sqrt{v^{2}(z)} which parameterise the relative position of the defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. An immediate consequence of this is that the combination

(z)2gμν(X1(z))\displaystyle\ell(z)^{-2}g_{\mu\nu}(X_{1}(z)) (62)

is a Weyl-invariant metric. However, it is only defined on the defect 𝒟1\mathcal{D}_{1}. If we could extend it away from 𝒟1\mathcal{D}_{1} and into the bulk MM, it would become a Weyl-invariant metric on MM and greatly simplify the problem of constructing Weyl-invariant effective actions. In this subsection we show that such an extension is indeed possible.

We first construct a canonical system of coordinates around 𝒟1\mathcal{D}_{1}, similar to Riemann normal coordinates. For this, we employ a Weyl-invariant equation for curves, analogous to the geodesic equation (15). No such second-order equation exists,131313Equation (19) is second-order, but contains derivatives of X1(z,t)X_{1}(z,t) with respect to zz. For reasons which will soon become apparent, here we want to construct a curve for an isolated value of zz. but infinitely-many can be found at third order.141414There are infinitely-many in a general curved space BaileyCircles . It is likely that it is unique in a conformally-flat space. We will use a particular version, as defined in BaileyCircles and called there the conformal circle equation,151515The conformal circle equation is not invariant under general changes of the parameter tt. The same is true for the geodesic equation (15): its solutions are affinely-parameterised geodesics. Similarly, the solutions to (63) are conformal circles with a parameterisation of a special type — a “projective” parameterisation BaileyCircles . In this context, an important property of (63) is that it is invariant under projective (fractional-linear) transformations of the parameter tt.

taμ=3(ua)u2aμ3a22u2uμ+u2uνPνμ2uαuβuμPαβ.\displaystyle\nabla_{t}a^{\mu}=\frac{3(u\cdot a)}{u^{2}}a^{\mu}-\frac{3a^{2}}{2u^{2}}u^{\mu}+u^{2}u^{\nu}P^{\mu}_{\nu}-2u^{\alpha}u^{\beta}u^{\mu}P_{\alpha\beta}. (63)

This is a third-order equation for a curve γμ(t)\gamma^{\mu}(t), where we used the notation uμ=tγμu^{\mu}=\nabla_{t}\gamma^{\mu}, aμ=t2γμa^{\mu}=\nabla_{t}^{2}\gamma^{\mu}, while PμνP_{\mu\nu} is the Schouten tensor defined in (266).

When MM is flat, the geodesics are straight lines, while the solutions to the conformal circle equation are straight lines and circles (see section 2.5). The appearance of the additional parameters — the radius of the circle and the plane in which it lies — corresponds to the fact that (63) is a third-order equation. In order to solve for γ\gamma, we need to specify not only the initial position γ(0)\gamma(0) and the initial velocity u(0)u(0), but also the initial covariant acceleration a(0)a(0). The latter transforms inhomogeneously under Weyl transformations as

δωaμ\displaystyle\delta_{\omega}a^{\mu} =2uμuννωu2μω.\displaystyle=2u^{\mu}u^{\nu}\partial_{\nu}\omega-u^{2}\partial^{\mu}\omega. (64)

This in particular means that simply setting a(0)=0a(0)=0 breaks Weyl invariance.

In general, there is no canonical way of specifying the initial condition for a(0)a(0). This makes it impossible to define canonical coordinates on a general conformal manifold by analogy to Riemann normal coordinates, which in turn makes the classification of general Weyl invariants a famously hard problem Fefferman:2007rka . Fortunately, it turns out that it is possible to choose a(0)a(0) canonically in the case at hand, where the conformal circle originates on one the two conformal defects 𝒟1,𝒟2\mathcal{D}_{1},\mathcal{D}_{2}. For now, we will assume that such a choice has been made; we describe it explicitly at the end of this section.

We can now construct coordinates in a neighbourhood of 𝒟1\mathcal{D}_{1} in the following way. Given a point xMx\in M close to 𝒟1\mathcal{D}_{1} we search for z𝒟1z\in\mathcal{D}_{1} and a vector yμy^{\mu} normal to 𝒟1\mathcal{D}_{1} at zz, such that the conformal circle with the initial condition γ(0)=X1(z)\gamma(0)=X_{1}(z), uμ(0)=yμu^{\mu}(0)=y^{\mu} (and the canonically chosen value for aμ(0)a^{\mu}(0)) passes through xx at t=1t=1, i.e. γ(1)=x\gamma(1)=x. We will show below that when xx is close enough to 𝒟1\mathcal{D}_{1}, the mapping x(z,y)x\mapsto(z,y) is well-defined, smooth, and one-to-one. We define the coordinates of xx to be the pair (z,y)(z,y), and denote the corresponding conformal circle by γx(t)\gamma_{x}(t).

This construction allows us to extend (z)\ell(z) to a function (x)\ell(x) of Weyl weight 1, defined on a neighbourhood of 𝒟1\mathcal{D}_{1}. Specifically, we define

2(x)γ˙x(0)2γ˙x2(1)2(z)=y2γ˙x2(1)2(z),\displaystyle\ell^{2}(x)\equiv\dot{\gamma}_{x}(0)^{-2}\dot{\gamma}_{x}^{2}(1)\ell^{2}(z)=y^{-2}\dot{\gamma}_{x}^{2}(1)\ell^{2}(z), (65)

where (z,y)(z,y) are the coordinates of xx, defined above. Since γ˙x(t)2\dot{\gamma}_{x}(t)^{2} is computed using the metric at γx(t)\gamma_{x}(t), this ensures that (x)\ell(x) transforms with Weyl weight 11 at xx, given that (z)\ell(z) transforms with Weyl weight 11 at X1(z)=γx(0)X_{1}(z)=\gamma_{x}(0). We will show below that (x)\ell(x) thus defined is in fact smooth and restricts to (z)\ell(z) on 𝒟1\mathcal{D}_{1}.

Using (x)\ell(x) we can now define the Weyl-invariant metric

g^μν(x)=2(x)gμν(x),\displaystyle\widehat{g}_{\mu\nu}(x)=\ell^{-2}(x)g_{\mu\nu}(x), (66)

which we refer to as the fusion metric. Any Weyl invariant that can be built out of gμνg_{\mu\nu} and vμv^{\mu} can now be written as simply a diffeomorphism invariant built out of g^μν\widehat{g}_{\mu\nu} and vμv^{\mu}. Indeed, since g^μν\widehat{g}_{\mu\nu} and gμνg_{\mu\nu} differ by a Weyl transformation, gμνg_{\mu\nu} can be replaced by g^μν\widehat{g}_{\mu\nu} in any Weyl invariant. Conversely, any diffeomorphism invariant of g^μν\widehat{g}_{\mu\nu} and vμv^{\mu} is automatically a Weyl invariant due to the Weyl invariance of g^μν\widehat{g}_{\mu\nu} and vμv^{\mu}.

We now give the details of the construction that were omitted above. To define the initial condition aμ(0)a^{\mu}(0) we simply observe that the vector

qμ=aμ(0)+2puμ(0)uσ(0)\RomanbarIIσ1pu2(0)\RomanbarIIμ+12aX1μu2(0)v2a(v2),\displaystyle q^{\mu}=a^{\mu}(0)+\frac{2}{p}u^{\mu}(0)u^{\sigma}(0)\Romanbar{II}_{\sigma}-\frac{1}{p}u^{2}(0)\Romanbar{II}^{\mu}+\frac{1}{2}\partial_{a}X_{1}^{\mu}u^{2}(0)v^{-2}\partial^{a}(v^{2}), (67)

where vμv^{\mu} parameterises the position of 𝒟2\mathcal{D}_{2} relative to 𝒟1\mathcal{D}_{1}, is a Weyl invariant. This is essentially the same calculation as verifying Weyl invariance of (19). Setting qμ=0q^{\mu}=0 defines a value for aμ(0)a^{\mu}(0) in a Weyl-invariant way, providing the necessary initial condition.

We now prove that (x)\ell(x) is smooth. The key subtlety here is that the conformal circle equation (63) contains u2u^{2} in the denominator. Because of this, it is not entirely obvious whether γ(1)\gamma(1) depends smoothly on u(0)u(0), and consequently whether xx depends smoothly on its coordinates (z,y)(z,y).

To solve this problem we first rewrite the conformal circle equation in terms of functions f,hf,h appearing in the potentially singular pieces,

f(ua)u2,ha2u2.\displaystyle f\equiv\frac{(u\cdot a)}{u^{2}},\qquad h\equiv\frac{a^{2}}{u^{2}}. (68)

For example, the conformal circle equation becomes

taμ=3faμ32huμ+u2uνPνμ2Pαβuαuβuμ.\displaystyle\nabla_{t}a^{\mu}=3fa^{\mu}-\tfrac{3}{2}hu^{\mu}+u^{2}u^{\nu}P^{\mu}_{\nu}-2P_{\alpha\beta}u^{\alpha}u^{\beta}u^{\mu}. (69)

It is easy to show (see appendix B) that tf\partial_{t}f and th\partial_{t}h can be expressed in a polynomial way in terms of u,a,f,hu,a,f,h. Overall, the system of equations for γ,u,a,f,h\gamma,u,a,f,h therefore has smooth functions in the right-hand side and thus the only possible singularities are in the initial conditions. Our initial conditions are given by γ(0)=X1(z)\gamma(0)=X_{1}(z), u(0)=yu(0)=y, and qμ=0q^{\mu}=0 (see (67)). This gives for f,hf,h

f(0)=1py\RomanbarII,h(0)=y2(1p2\RomanbarII2+14v4a(v2)a(v2)).\displaystyle f(0)=-\frac{1}{p}y\cdot\Romanbar{II},\quad h(0)=y^{2}\left(\frac{1}{p^{2}}\Romanbar{II}^{2}+\frac{1}{4}v^{-4}\partial_{a}(v^{2})\partial^{a}(v^{2})\right). (70)

We therefore find that the initial conditions for γ,u,a,f,h\gamma,u,a,f,h depend smoothly on zz and yy, and thus the solution γ(1)\gamma(1) depends smoothly on zz and yy as well. It is easy to verify that the differential of the map (z,y)γ(1)(z,y)\mapsto\gamma(1) is non-degenerate at y=0y=0, which shows that it is locally one-to-one and has a smooth inverse.

Finally, we note that

f(t)=12tlogu2(t).\displaystyle f(t)=\tfrac{1}{2}\partial_{t}\log u^{2}(t). (71)

Therefore,

γ˙2(1)γ˙2(0)=e201𝑑tf(t).\displaystyle\dot{\gamma}^{2}(1)\dot{\gamma}^{-2}(0)=e^{2\int_{0}^{1}dtf(t)}. (72)

Since, as discussed above, f(t)f(t) depends on zz and yy in a smooth way, it follows that γ˙2(1)γ˙2(0)\dot{\gamma}^{2}(1)\dot{\gamma}^{-2}(0) is also a smooth function of z,yz,y, and thus of xx. This shows that (65) depends smoothly on xx. Note, however, that due to the v4v^{-4} factor in (70), (x)\ell(x) is not analytic in vμv^{\mu}. It is, however, analytic in the derivatives of vμv^{\mu}, which is what will be important for our applications.

2.4 Properties of the fusion metric

In the previous subsection we described an procedure which constructs the fusion metric g^\widehat{g} given as initial data the physical metric gg in the bulk and the vector field vv on 𝒟1\mathcal{D}_{1}.161616The result only depends on gμνvμvν\sqrt{g_{\mu\nu}v^{\mu}v^{\nu}}, but talking about vv is easier since it is Weyl-invariant which saves us from discussing extraneous details. In this section show that g^\widehat{g} is not arbitrary and satisfies non-trivial local identities such as (82) and (84) regardless of the choice of gg. These identities will be important when we write down the independent local invariants of g^\widehat{g} in the following sections. Finally, we make some comments about isometries of g^\widehat{g}.

We will treat vv as fixed (together with the bulk manifold MM and other data) and view g^\widehat{g} as a function of gg. We will express this as g^=𝒢F(g)\widehat{g}=\mathcal{G}_{F}(g), where

𝒢F(g)=2g,\displaystyle\mathcal{G}_{F}(g)=\ell^{-2}g, (73)

and (x)\ell(x) depends on gg and is defined in section 2.3. Our goal is to determine which metrics are in the image of 𝒢F\mathcal{G}_{F}.

It is easy to see that 𝒢F\mathcal{G}_{F} is not surjective – there are many metrics which cannot be represented as 𝒢F(g)\mathcal{G}_{F}(g) for any gg. Indeed, by construction 𝒢F\mathcal{G}_{F} is Weyl-invariant,

𝒢F(e2ωg)=𝒢F(g).\displaystyle\mathcal{G}_{F}(e^{2\omega}g)=\mathcal{G}_{F}(g). (74)

Furthermore, by (73) 𝒢F(g)\mathcal{G}_{F}(g) is Weyl-equivalent to gg. This implies that the image of 𝒢F\mathcal{G}_{F} contains precisely one metric from each conformal class. Such metrics are easy to characterise formally; the above identities imply that if g^=𝒢F(g)\widehat{g}=\mathcal{G}_{F}(g) for some gg, then

𝒢F(g^)=g^.\displaystyle\mathcal{G}_{F}(\widehat{g})=\widehat{g}. (75)

An important property of 𝒢F(g)\mathcal{G}_{F}(g) is that its Taylor coefficients around 𝒟1\mathcal{D}_{1} depend locally on gg. Therefore, the constraint (75) can be rewritten as a set of constraints on the Taylor coefficients of g^\widehat{g}. In the rest of this section we describe a simple way to derive these constraints.

Let ^(x)\widehat{\ell}(x) be the function \ell computed for a metric g^\widehat{g} which satisfies (75). Equation (75) together with (73) imply that ^(x)=1\widehat{\ell}(x)=1. It then follows from the definition (65) that the curves γ^x(t)\widehat{\gamma}_{x}(t) constructed for the fusion metric satisfy

γ^˙x2(1)=γ^˙x2(0).\displaystyle\dot{\widehat{\gamma}}_{x}^{2}(1)=\dot{\widehat{\gamma}}_{x}^{2}(0). (76)

It is easy to check that γ^γ^x(s)(t)=γ^x(su)\widehat{\gamma}_{\widehat{\gamma}_{x}(s)}(t)=\widehat{\gamma}_{x}(su) from the definition of γ^x\widehat{\gamma}_{x}, and therefore

γ^˙x2(st)=s2γ^˙γ^x(s)2(t).\displaystyle\dot{\widehat{\gamma}}_{x}^{2}(st)=s^{-2}\dot{\widehat{\gamma}}_{\widehat{\gamma}_{x}(s)}^{2}(t). (77)

Together with (76), this implies

γ^˙x2(t)=γ^˙x2(0).\displaystyle\dot{\widehat{\gamma}}_{x}^{2}(t)=\dot{\widehat{\gamma}}_{x}^{2}(0). (78)

Conversely, if some metric g^\widehat{g} satisfies (78) for all curves γ^x\widehat{\gamma}_{x} and ^(z)=vv=1\widehat{\ell}(z)=\sqrt{v\cdot v}=1, then (65) implies ^(x)=1\widehat{\ell}(x)=1 and thus g^\widehat{g} satisfies (75).

The constraint (78) can be equivalently replaced by local constraints, which together with ^(z)=vv=1\widehat{\ell}(z)=\sqrt{v\cdot v}=1 read

{tnγ^˙x2(t)|t=0=0,n>0v2=1.\displaystyle\begin{cases}\partial_{t}^{n}\dot{\widehat{\gamma}}_{x}^{2}(t)|_{t=0}=0,&\forall n>0\\ v^{2}=1.\end{cases} (79)

The above discussion implies that these conditions are equivalent to (75). The first few of these constraints are

v2=1,a(0)u(0)=0,^ta(0)u(0)+a(0)a(0)=0,,\displaystyle v^{2}=1,\quad a(0)\cdot u(0)=0,\quad\widehat{\nabla}_{t}a(0)\cdot u(0)+a(0)\cdot a(0)=0,\quad\cdots, (80)

where as usual u=^tγ^x,a=^t2γ^xu=\widehat{\nabla}_{t}\widehat{\gamma}_{x},\,a=\widehat{\nabla}^{2}_{t}\widehat{\gamma}_{x}. In this formulation we can choose u(0)u(0) and γx(0)\gamma_{x}(0) at will, since these quantities are the initial conditions for the curves γ^x\widehat{\gamma}_{x}.

The constraint v2=g^μνvμvν=1v^{2}=\widehat{g}_{\mu\nu}v^{\mu}v^{\nu}=1 is expressed directly in terms of g^\widehat{g}. On the other hand, the subsequent constraints such as a(0)u(0)=0a(0)\cdot u(0)=0 are more implicit. Using the initial condition qμ=0q^{\mu}=0 and (67), we find that a(0)u(0)=0a(0)\cdot u(0)=0 is equivalent to

u2(0)uσ(0)\RomanbarII^σ=0.\displaystyle u^{2}(0)u^{\sigma}(0)\widehat{\Romanbar{II}}_{\sigma}=0. (81)

This has to be satisfied for all u(0)u(0) and for all points on 𝒟1\mathcal{D}_{1}. This implies

\RomanbarII^μ=0.\displaystyle\widehat{\Romanbar{II}}^{\mu}=0. (82)

It then also follows from q=0q=0, v2=1v^{2}=1 and (67) that a(0)=0a(0)=0. Note that in the case p=1p=1, \RomanbarII^μ=0\widehat{\Romanbar{II}}^{\mu}=0 is equivalent to 𝒟1\mathcal{D}_{1} being a geodesic for g^\widehat{g}.

Consider now the condition ^ta(0)u(0)+a(0)a(0)=0\widehat{\nabla}_{t}a(0)\cdot u(0)+a(0)\cdot a(0)=0. Using a(0)=0a(0)=0 and the conformal circle equation (63) we find

0=u2(0)uμ(0)uν(0)P^μν.\displaystyle 0=u^{2}(0)u^{\mu}(0)u^{\nu}(0)\widehat{P}_{\mu\nu}. (83)

This has to be satisfied for all u(0)u(0), which implies that all normal components of the Schouten tensor P^μν\widehat{P}_{\mu\nu} vanish,

P^μν=0,μ,ν normal.\displaystyle\widehat{P}_{\mu\nu}=0,\quad\mu,\nu\text{ normal}. (84)

Note that the number of constraints v2=1v^{2}=1, (82), and (84) is precisely the same as the number of the coefficients in the Taylor expansion (in normal directions) of a scalar function up to the second order. This is expected since we are effectively writing out the equation ^(x)=1\widehat{\ell}(x)=1 order-by-order. Higher-order constraints can be derived in the same way by taking t\nabla_{t} derivatives of the conformal circle equation (63), and using the resulting equation in (80). They will involve higher-derivative curvature invariants.

This discussion allows us to give an alternative characterisation of the fusion metric g^\widehat{g} for a given physical metric gg: to the quadratic order in the distance from 𝒟1\mathcal{D}_{1}, the fusion metric g^\widehat{g} is the unique metric in the conformal class of gg which satisfies g^μνvμvν=0\widehat{g}_{\mu\nu}v^{\mu}v^{\nu}=0, (82), and (84). The scale factor (x)\ell(x) is then defined by g^=2g\widehat{g}=\ell^{-2}g. Existence and uniqueness are easy to check from the Weyl transformation laws of \RomanbarII^μ\widehat{\Romanbar{II}}^{\mu} and P^μν\widehat{P}_{\mu\nu}. If we classify the analogous constraints at all orders, they will provide an all-order characterisation of g^\widehat{g}.

This characterisation is clearly not unique starting at the second order. Specifically, we could require that the purely normal components of R^μν\widehat{R}_{\mu\nu} vanish, or we could use any other purely normal symmetric rank-2 2-derivative tensor. Similar ambiguities exist at higher orders. Of course, one would have to verify that a metric solving such an alternative condition exists. The advantage of the construction in section 2.3 is that it is guaranteed to work at all orders. The minor drawback is that the constraints such as (84) have to be derived order-by-order rather than chosen at will.

Isometries of the fusion metric

Finally, we note that any conformal symmetry of the physical metric gg which preserves 𝒟1\mathcal{D}_{1} and does not modify171717Note that here we refer to the value of \ell no the defect. (z)\ell(z) will become an isometry of the fusion metric g^\widehat{g}. This is because g^\widehat{g} depends only on (z)\ell(z) and the conformal class of gg. For example, transverse rotations around 𝒟1\mathcal{D}_{1} always preserve (z)\ell(z) (regardless of its value), so whenever they are available, they become isometries of g^\widehat{g}. For example, this always applies for spherical defects in flat space.

2.5 Example: spherical defects in flat space II

Let us now consider how the construction of section 2.3 works in the case of flat space and spherical defects.

First of all, the flat space version of the conformal circle equation (63) becomes

taμ=3(ua)u2aμ3a22u2uμ.\displaystyle\partial_{t}a^{\mu}=\frac{3(u\cdot a)}{u^{2}}a^{\mu}-\frac{3a^{2}}{2u^{2}}u^{\mu}. (85)

Let us study its solutions. Using translations, we can assume that γμ(0)=0\gamma^{\mu}(0)=0. The remaining initial conditions are given by the vectors aμ(0)a^{\mu}(0) and uμ(0)u^{\mu}(0). Using rotations, we can make sure that these vectors lie in the 1-2 plane. Then the whole curve γμ(t)\gamma^{\mu}(t) will be contained in this plane, and thus we can fully characterise the solution by the function

s(t)=γ1(t)+iγ2(t).\displaystyle s(t)=\gamma^{1}(t)+i\gamma^{2}(t). (86)

It is easy to check that in terms of s(t)s(t) the conformal circle equation becomes simply

{s(t),t}s˙˙˙(t)s˙(t)32(s¨(t)s˙(t))2=0,\displaystyle\{s(t),t\}\equiv\frac{\dddot{s}(t)}{\dot{s}(t)}-\frac{3}{2}\left(\frac{\ddot{s}(t)}{\dot{s}(t)}\right)^{2}=0, (87)

where {s(t),t}\{s(t),t\} is the Schwarzian derivative of s(t)s(t). The only functions with vanishing Schwarzian derivative are the fractional-linear transformations, i.e. functions of the form

s(t)=c11t+c12c21t+c22,\displaystyle s(t)=\frac{c_{11}t+c_{12}}{c_{21}t+c_{22}}, (88)

for some cijc_{ij}\in\mathbb{C}. Since tt takes values in \mathbb{R}, and the image of \mathbb{R} under a fractional-linear map is a circle, it follows that the curve γ\gamma follows a circle (possibly of infinite radius, i.e. a line).

To express the solution in terms of initial data, define

α=s¨(0)s˙(0)2,\displaystyle\alpha=\frac{\ddot{s}(0)}{\dot{s}(0)^{2}}, (89)

and recall s(0)=γ1(0)+iγ2(0)=0s(0)=\gamma^{1}(0)+i\gamma^{2}(0)=0. The solution takes the form

s(t)=s˙(0)t1αs˙(0)t2.\displaystyle s(t)=\frac{\dot{s}(0)t}{1-\frac{\alpha\dot{s}(0)t}{2}}. (90)

This explicit form implies that

s()=2α,s˙(t)=4α2s˙(0)t2+O(t3).\displaystyle s(\infty)=-\frac{2}{\alpha},\quad\dot{s}(t)=\frac{4}{\alpha^{2}\dot{s}(0)t^{2}}+O(t^{-3}). (91)

These identities will be useful later.

Let us apply the above discussion to the curves γx\gamma_{x} defined in section 2.3. Firstly, the invariant qq defined by (67) becomes, in flat space and for a flat defect 𝒟1\mathcal{D}_{1},

qμ=aμ(0)+12aX1μu2(0)v2av2.\displaystyle q^{\mu}=a^{\mu}(0)+\frac{1}{2}\partial_{a}X_{1}^{\mu}u^{2}(0)v^{-2}\partial^{a}v^{2}. (92)

Setting q=0q=0 then fixes

aμ(0)u2(0)=12aX1μv2av2=12aX1μ2a2.\displaystyle\frac{a^{\mu}(0)}{u^{2}(0)}=-\frac{1}{2}\partial_{a}X_{1}^{\mu}v^{-2}\partial^{a}v^{2}=-\frac{1}{2}\partial_{a}X_{1}^{\mu}\ell^{-2}\partial^{a}\ell^{2}. (93)

Note that this in particular says that aμ(0)a^{\mu}(0) is tangent to the defect. On the other hand, the initial velocity uμ(0)u^{\mu}(0) is, by construction, orthogonal to the defect.

Since a(0)u(0)=0a(0)\cdot u(0)=0, by choosing the coordinates appropriately, s˙(0)\dot{s}(0) can be taken to real and positive, while s¨(0)\ddot{s}(0) can be taken to be imaginary with positive imaginary part, so that α\alpha is imaginary. In this case the real axis is normal to the defect, while the imaginary axis is along the defect. Using (91) we then see that s()s(\infty) is imaginary and thus the conformal circle γ(t)\gamma(t) meets the flat defect at t=t=\infty. The conformal circle is normal to the defect at that point, since s˙(t)\dot{s}(t) is real to the leading order according to (91).

Furthermore, note that the position of γ()\gamma(\infty) depends only on α\alpha, i.e. the vector aμ(0)/u(0)2a^{\mu}(0)/u(0)^{2}. This vector is fixed by (93) and is independent of uμ(0)u^{\mu}(0). This shows that given (z)=v2(z)\ell(z)=\sqrt{v^{2}(z)}, for every point zz on 𝒟1\mathcal{D}_{1} there is a corresponding point r(z)r_{\ell}(z) defined by γ(0)=z\gamma(0)=z and r(z)=γ()r_{\ell}(z)=\gamma(\infty), and depending only on zz and \ell. For different choices of uμ(0)u^{\mu}(0), the corresponding conformal circles all pass through zz at t=0t=0 and r(z)r_{\ell}(z) at t=t=\infty, being normal to 𝒟1\mathcal{D}_{1} at these intersections.

While the above statements have been derived for a flat defect, the are conformally-invariant and therefore apply to any spherical defect. To determine the bulk length scale (x)\ell(x) at a given point xx in the bulk, one has to first determine on which conformal circle xx lies, parameterise the circle appropriately, and then apply (65). These steps can in principle be performed without solving any differential equations by using the above statements.

In practice, it appears to be hard to carry out this procedure for a general defect length scale (z)\ell(z) or even for the (z)\ell(z) which arise from the displacements described in section 2.2. However, some simple observations can be made in full generality.

Spherical defects in flat space always admit transverse rotations. As discussed in section 2.4, in the fusion metric transverse rotations become isometries. This places strong constrains on tensors with normal indices – only their singlet components can be non-zero. For example, for spherical defects we always have

\RomanbarII^abμ=0,\displaystyle\widehat{\Romanbar{II}}_{ab}^{\mu}=0, (94)

which is a strengthening of (82). Similarly, condition (84) gets strengthened to

P^μν=0,(if μ or ν is normal).\displaystyle\widehat{P}_{\mu\nu}=0,\quad(\text{if $\mu$ or $\nu$ is normal}). (95)

Equation (94) and the Gauss equation (270) imply that the intrinsic Riemann tensor coincides with the restriction of the bulk Riemann tensor for the fusion metric. These and other constraints serve to reduce the number of independent couplings in the fusion effective action for spherical defects in flat space.

In the rest of this section we will explicitly work out the fusion metric in several examples where (z)\ell(z) is constant.

Example: line defects in d=3d=3

We first consider the case of circular line defects in d=3d=3, in situations when (z)\ell(z) is constant. Without loss of generality, we can assume that 𝒟1\mathcal{D}_{1} is a circle of radius 11 centred around 0 and lying in the 1-2 plane, x3=0x^{3}=0. Since (z)=\ell(z)=\ell is constant, the map zr(z)z\to r_{\ell}(z) has to respect the Euclidean symmetries of this circle, which include reflections in planes normal to a diameter of 𝒟1\mathcal{D}_{1}. The only option is

r(z)=z,\displaystyle r_{\ell}(z)=-z, (96)

where z-z denotes the point on 𝒟1\mathcal{D}_{1} diametrically opposite to zz.

Consider now a bulk point x0x_{0} and suppose that it is in the 1-3 plane, x02=0x^{2}_{0}=0. We will be able to recover the general case using rotations around the axis of 𝒟1\mathcal{D}_{1}. We need to work in a neighbourhood of 𝒟1\mathcal{D}_{1}, so we assume that (x01,x03)(x^{1}_{0},x^{3}_{0}) is close to (1,0)(1,0).

The conformal circle which passes through x0x_{0} has to intersect 𝒟1\mathcal{D}_{1} at x2=0x^{2}=0 and x1=±1x^{1}=\pm 1. Using the complex coordinate s=x1+ix3s=x^{1}+ix^{3}, the conformal circles which pass through s=1s=1 at t=0t=0 and through s=1s=-1 at t=t=\infty are given by

s(t)=1+s˙(0)t/21s˙(0)t/2.\displaystyle s(t)=\frac{1+\dot{s}(0)t/2}{1-\dot{s}(0)t/2}. (97)

Requiring that s(1)=s0=x01+ix03s(1)=s_{0}=x_{0}^{1}+ix_{0}^{3} we find

s˙(0)=2s01s0+1,s˙(1)=s0212.\displaystyle\dot{s}(0)=2\frac{s_{0}-1}{s_{0}+1},\quad\dot{s}(1)=\frac{s_{0}^{2}-1}{2}. (98)

This gives

(x0)=|s˙(1)||s˙(0)|=|s0+1|24.\displaystyle\ell(x_{0})=\frac{|\dot{s}(1)|}{|\dot{s}(0)|}\ell=\frac{|s_{0}+1|^{2}}{4}\ell. (99)

If we now return to the case of general x0x_{0} and parameterise it using the cylindrical coordinates (r,ϕ,x3)(r,\phi,x^{3}) with x1=rcosϕx^{1}=r\cos\phi, x2=rsinϕx^{2}=r\sin\phi, we get

(r,ϕ,x3)=(x3)2+(r+1)24,\displaystyle\ell(r,\phi,x^{3})=\frac{(x^{3})^{2}+(r+1)^{2}}{4}\ell, (100)

which is smooth in the neighbourhood of 𝒟1\mathcal{D}_{1} given by r1,x30r\approx 1,x_{3}\approx 0 but has a singularity on the axis r=0r=0.

This metric best interpreted in the toroidal coordinates τ,σ,ϕ\tau,\sigma,\phi defined by τ0\tau\geq 0, σ[π,π]\sigma\in[-\pi,\pi] and

r\displaystyle r =sinhτcoshτcosσ,x3=sinσcoshτcosσ.\displaystyle=\frac{\sinh\tau}{\cosh\tau-\cos\sigma},\quad x^{3}=\frac{\sin\sigma}{\cosh\tau-\cos\sigma}. (101)

In these coordinates we have

(τ,σ,ϕ)=eτ2(coshτcosσ).\displaystyle\ell(\tau,\sigma,\phi)=\frac{e^{\tau}\ell}{2(\cosh\tau-\cos\sigma)}. (102)

The physical flat-space metric is

ds2=(coshτcosσ)2(dτ2+dσ2+sinh2τdϕ2),\displaystyle ds^{2}=(\cosh\tau-\cos\sigma)^{-2}\left(d\tau^{2}+d\sigma^{2}+\sinh^{2}\tau d\phi^{2}\right), (103)

and the fusion metric is

d^s2=2(τ,σ,ϕ)ds2=42e2τ(dτ2+dσ2+sinh2τdϕ2).\displaystyle\widehat{d}s^{2}=\ell^{-2}(\tau,\sigma,\phi)ds^{2}=4\ell^{-2}e^{-2\tau}\left(d\tau^{2}+d\sigma^{2}+\sinh^{2}\tau d\phi^{2}\right). (104)

This metric is invariant under translations in σ\sigma and ϕ\phi. Translations in σ\sigma are new as compared to the physical metric. They coincide with transverse rotations around 𝒟1\mathcal{D}_{1}, in agreement with our general expectations. In r,ϕ,x3r,\phi,x^{3} coordinates the only non-zero component of the Schouten tensor is

P^ϕϕ=2r(1+r2)+(x3)2.\displaystyle\widehat{P}_{\phi\phi}=\frac{2r}{(1+r^{2})+(x^{3})^{2}}. (105)

This obviously satisfies the condition (95).

Example: parallel dimension-pp flat defects

This is a continuation of an example in section 2.2. As in section 2.2, we shall place 𝒟1\mathcal{D}_{1} so that it passes through x=0x=0, and spans the coordinate directions 1,2,,p1,2,\cdots,p. By translation and rotational symmetry of the defect, we can see that r(z)=r_{\ell}(z)=\infty.

If we consider a point x0x_{0} in the bulk, then without loss of generality, we can assume that x0=ce^dx_{0}=c\widehat{e}_{d}, as any other point can be recovered by transverse rotations and translations along the defect. In terms of s(t)s(t), r(z)=s()=r_{\ell}(z)=s(\infty)=\infty together with (91) implies α=0\alpha=0. Equation (90) then implies that

s(t)=s˙(0)t,\displaystyle s(t)=\dot{s}(0)t, (106)

and thus γμ\gamma^{\mu} has constant velocity. Equation (65) then gives us that (x0)=(0)\ell(x_{0})=\ell(0). By symmetry under transverse rotations, and translations along the defect, this must hold for any x0x_{0}, and so (x)\ell(x) is a constant everywhere in the bulk.

We saw in (57) that if we have a second parallel defect at a distance LL away, then (z)=L\ell(z)=L on the defect, and so we must also have that

(x)=L\displaystyle\ell(x)=L (107)

in the bulk as well. The fusion metric is the flat ds^2=L2ds2d\widehat{s}^{2}=L^{-2}ds^{2}.

Example: parallel flat line defects in ×Sd1\mathbb{R}\times S^{d-1}

In order to study a pair of parallel flat line defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} in ×Sd1\mathbb{R}\times S^{d-1}, we shall again use the exponential map (58) to map the defects to rays in {0}\mathbb{R}\setminus\{0\} starting at zero and pointing in directions Ω1\Omega_{1} and Ω2\Omega_{2} respectively. We then take a bulk point, which in the {0}\mathbb{R}\setminus\{0\} picture is located at a point x=eτ0Ω0x=e^{\tau_{0}}\Omega_{0} for some τ0\tau_{0}\in\mathbb{R}, Ω0Sd1\Omega_{0}\in S^{d-1}.

Without loss of generality, we take

Ω1=(1,0,0,,0),Ω0=(cosθ,sinθ,0,0,0).\displaystyle\Omega_{1}=(1,0,0,\dots,0),\qquad\Omega_{0}=(\cos\theta,\sin\theta,0,0,\dots 0). (108)

By reflection symmetries γ\gamma must lie in the 121-2 plane, and by inversion symmetry under τ2τ0τ\tau\to 2\tau_{0}-\tau it must lie along the circle going through xx and z=eτ0Ω1z=e^{\tau_{0}}\Omega_{1}. The SO(d1)\mathrm{SO}(d-1) rotations which preserve Ω1\Omega_{1} imply that r(z)=zr_{\ell}(z)=-z.181818Though this point is not on the defect, it is on the straight line obtained by extending 𝒟1\mathcal{D}_{1} through the origin. Letting s(t)=γ1(t)+iγ2(t)s(t)=\gamma^{1}(t)+i\gamma^{2}(t), z~=eτ0\widetilde{z}=e^{\tau_{0}}, x~=eτ0+iθ\widetilde{x}=e^{\tau_{0}+i\theta} this implies

s(t)=z~t(z~x~)+z~(z~+x~)t(z~x~)+(z~+x~).\displaystyle s(t)=\frac{-\widetilde{z}t(\widetilde{z}-\widetilde{x})+\widetilde{z}(\widetilde{z}+\widetilde{x})}{t(\widetilde{z}-\widetilde{x})+(\widetilde{z}+\widetilde{x})}. (109)

We can now use this expression in the conformal frame of ×Sd1\mathbb{R}\times S^{d-1} to find

(τ,θ)=αcos2θ2.\displaystyle\ell(\tau,\theta)=\alpha\cos^{2}\frac{\theta}{2}. (110)

This result holds in generality where α\alpha is the angle on the cylinder between 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}, and θ\theta is the angle between 𝒟1\mathcal{D}_{1} and xx.

A useful set of coordinates on Sd1S^{d-1} are the stereographic coordinates uiu^{i} for i=2di=2\dots d

ui=2Ωi1+Ω1.\displaystyle u^{i}=\frac{2\Omega^{i}}{1+\Omega^{1}}. (111)

Note that the defect 𝒟1\mathcal{D}_{1} is at u=0u=0. The fusion metric can be written as

ds^2\displaystyle d\widehat{s}^{2} =α2i=2ddui2+α2(1+u24)2dτ2.\displaystyle={\alpha^{-2}}\sum_{i=2}^{d}du_{i}^{2}+\alpha^{-2}{(1+\tfrac{u^{2}}{4})^{2}}d\tau^{2}. (112)

The only non-zero component of the corresponding Schouten tensor is

P^ττ\displaystyle\widehat{P}_{\tau\tau} =12(1+u24),\displaystyle=-\frac{1}{2}\left(1+\frac{u^{2}}{4}\right), (113)

and the Ricci scalar is given by

R^\displaystyle\widehat{R} =2(d1)g^ττP^ττ=(d1)α21+u24.\displaystyle=2(d-1)\widehat{g}^{\tau\tau}\widehat{P}_{\tau\tau}=-\frac{(d-1)\alpha^{2}}{1+\frac{u^{2}}{4}}. (114)

Note for future reference that only the even-order derivatives of the fusion metric (112) are non-zero on the defect at u=0u=0.

3 Effective action

In this section we study the effective action for the fusion of two defects. We use the same conventions as in section 2. In this section, γ\gamma always refers to the defect metric.

3.1 General comments

Recall that our working assumption is that the fusion of two conformal defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} can be viewed as an RG flow terminating at an IR conformal defect 𝒟Σ\mathcal{D}_{\Sigma}. Explicitly, in terms of correlation functions we expect

𝒟1𝒟2𝒟Σ[eSeff],\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\cdots\rangle\approx\langle\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}}]\cdots\rangle, (115)

where \cdots represent other insertions. Here, the characteristic distance LL between the two defects goes to zero, L0L\to 0, while all the other distances remain finite on the order of the IR scale RR. The effective action SeffS_{\text{eff}} describes the end tail of the RG flow and has the general form

Seff=dpzγ(λ1𝟏+𝒪λ𝒪𝒪),\displaystyle S_{\text{eff}}=\int d^{p}z\sqrt{\gamma}\left(\lambda_{1}\mathbf{1}+\sum_{{\mathcal{O}}}\lambda_{\mathcal{O}}{\mathcal{O}}\right), (116)

where λ𝒪\lambda_{\mathcal{O}} are various couplings. Here, we include the identity operator and a sum over (marginally) irrelevant operators primary 𝒪{\mathcal{O}} that exist on 𝒟Σ\mathcal{D}_{\Sigma}. Descendant couplings can be omitted because they can be reduced to primary couplings by integration by parts. No (marginally) relevant operators are included because their existence would contradict the assumption that we reach the IR fixed point 𝒟Σ\mathcal{D}_{\Sigma}, see section 3.8. Finally, we do not include O(1)O(1) 0-derivative couplings for the exactly marginal operators. Instead, we use the convention where these are absorbed by choosing for 𝒟Σ\mathcal{D}_{\Sigma} an appropriate point in the conformal manifold. Thus, marginal couplings start at 1-derivative order or O(L)O(L).

That the effective action (116) has to be an integral of a local Lagrangian density follows from the locality of the defects 𝒟1,𝒟2\mathcal{D}_{1},\mathcal{D}_{2}, and 𝒟Σ\mathcal{D}_{\Sigma}. Correspondingly, the couplings λ𝒪\lambda_{\mathcal{O}} have to be local functionals of the various parameters present in the correlation function. The latter are the usual geometric quantities such as the bulk and defect metrics as well as the extrinsic defect geometry for 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} and their relative position.

In practice, it is useful to replace all the geometric information about 𝒟2\mathcal{D}_{2} by the displacement vector field vμ(z)v^{\mu}(z) on 𝒟1\mathcal{D}_{1}, defined in section 2.1, that describes the position of 𝒟2\mathcal{D}_{2} relative to 𝒟1\mathcal{D}_{1}. Furthermore, we can assume that 𝒟Σ\mathcal{D}_{\Sigma} is inserted along the same submanifold as 𝒟1\mathcal{D}_{1} — as we discuss below, any changes to this can be absorbed into the coupling λD\lambda_{D} for the displacement operator. Therefore, the couplings λ𝒪\lambda_{\mathcal{O}} have to be constructed from the bulk geometry, the geometry of 𝒟1\mathcal{D}_{1}, and the vector field vμ(z)v^{\mu}(z).

Dimensional analysis implies that λ𝒪LΔ𝒪pf(L/R)\lambda_{\mathcal{O}}\sim L^{\Delta_{\mathcal{O}}-p}f(L/R), where RR is the IR scale. On the other hand, the dependence on RR can only appear as the powers of 1/R1/R coming from derivatives of the metric and other parameters. Therefore, as is usual, the derivative expansion of the couplings λ𝒪\lambda_{\mathcal{O}} becomes an expansion in powers of L/RL/R. If we are only interested in approximating the correlation function in the left-hand side of (115) to order O(La)O(L^{a}), we can truncate the effective action (116) to only include terms with Δ𝒪+npa\Delta_{\mathcal{O}}+n-p\leq a, where nn is number of derivatives in λ𝒪\lambda_{\mathcal{O}}, and evaluate the conformal perturbation theory to O(La)O(L^{a}). Notice that the only coupling that does not go to zero in the limit L0L\to 0 is the identity operator coupling λ1\lambda_{1}.

Before proceeding with the more detailed analysis of the effective action, let us comment on the “normalisation” of the conformal defects 𝒟1,𝒟2\mathcal{D}_{1},\mathcal{D}_{2} and 𝒟Σ\mathcal{D}_{\Sigma}. Throughout the paper, we assume that these defects are normalised so that they are both local and Weyl-invariant up to Weyl anomalies. Note that changing the defect normalisation 𝒟λ𝒟\mathcal{D}\to\lambda\mathcal{D} for λ\lambda\in\mathbb{C} is not a local modification of the defect action and thus breaks locality. In other words, the only normalisation changes that we allow are the scheme changes of the form

𝒟edpzγN𝒟,\displaystyle\mathcal{D}\to e^{\int d^{p}z\sqrt{\gamma}N}\mathcal{D}, (117)

where NN is a mass dimension pp diffeomorphism-invariant local quantity. In particular, it is impossible in general to set 𝒟=1\langle\mathcal{D}\rangle=1, which is a normalisation sometimes used in the literature. Note that unless γN\sqrt{\gamma}N is Weyl-invariant, this change of scheme will change the form of the Weyl anomaly. We do not generally commit to a particular form of the Weyl anomaly.

If the action of 𝒟1\mathcal{D}_{1}, 𝒟2\mathcal{D}_{2}, or 𝒟Σ\mathcal{D}_{\Sigma} is modified using (117), this may change the Wilson coefficients in the effective action SeffS_{\text{eff}}. We do not consider such modifications of the effective action and treat the actions of the defects 𝒟1,𝒟2\mathcal{D}_{1},\mathcal{D}_{2}, and 𝒟Σ\mathcal{D}_{\Sigma} as fixed.

Conformal boundaries

A special case of our discussion is when 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are conformal boundaries, and the bulk CFT is sandwiched in between them. In this case, no bulk is left in the L0L\to 0 limit. The IR defect 𝒟Σ\mathcal{D}_{\Sigma} becomes a (d1)(d-1)-dimensional theory. Generically, one expects 𝒟Σ\mathcal{D}_{\Sigma} to be gapped, in which case only the identity operator appears in SeffS_{\text{eff}}. If 𝒟Σ\mathcal{D}_{\Sigma} is a CFT, our general discussion still applies.

3.2 Parity and reality conditions

Both the bulk CFT and the defects 𝒟1,𝒟2\mathcal{D}_{1},\mathcal{D}_{2} can preserve or break space parity. In this section we briefly discuss the possible breaking patterns on general manifolds MM and the implications for reality conditions on the effective couplings λ𝒪\lambda_{\mathcal{O}}.

If the bulk CFT preserves parity, this means that its partition function on a general manifold MM can be computed without specifying an orientation of MM.191919A parity-preserving theory may have parity-odd local operators, which require specification of an orientation at the point of their insertion. In particular, it can be studied on non-orientable manifolds. If the bulk CFT breaks parity, we need to specify an orientation of MM in order to compute the partition function. In any case, we will assume that the bulk theory satisfies the following Hermiticity axiom Atiyah:1989vu ; Kontsevich:2021dmb ,

𝒵(M)¯=𝒵(M¯),\displaystyle\overline{\mathcal{Z}(M)}=\mathcal{Z}(\overline{M}), (118)

where 𝒵(M)\mathcal{Z}(M) denotes the partition function on a manifold MM and M¯\overline{M} denotes MM with its orientation reversed. In particular, parity-preserving theories have real partition functions. This identity allows MM to have a boundary, in which case 𝒵(M)\mathcal{Z}(M) is valued in an appropriate vector space and 𝒵(M¯)\mathcal{Z}(\overline{M}) in the complex conjugate vector space. This Hermiticity axiom (or a somewhat restricted version thereof) is necessary to be able to formulate reflection positivity.

A conformal defect can have a richer parity breaking structure. If the bulk is parity-preserving, the defect can break reflections along the defect, transverse to the defect, or both. For example, Wilson lines in complex representations break parity along the defect, while monodromy defects generally break (at least the naive) transverse reflections. On a general manifold MM this means that in order to compute partition functions involving a defect 𝒟\mathcal{D}, we may need to specify an orientation on 𝒟\mathcal{D}, as well as an orientation in the normal bundle of 𝒟\mathcal{D}.

If the bulk CFT breaks parity, these two orientations are not independent since they can be related by the bulk orientation. By convention, we can agree that, if any orientation needs to be specified for the defect, it is only the defect orientation. Thus, if the bulk CFT breaks parity, the defect still may break or preserve parity along the defect. Preserving defect parity while breaking bulk parity is possible because a defect reflection can be achieved by a bulk rotation.

In the presence of a defect the Hermiticity axiom needs to be modified as follows,

𝒵(M,𝒟)¯=𝒵(M¯,𝒟¯),\displaystyle\overline{\mathcal{Z}(M,\mathcal{D})}=\mathcal{Z}(\overline{M},\overline{\mathcal{D}}), (119)

where 𝒟¯\overline{\mathcal{D}} denotes 𝒟\mathcal{D} with the defect orientation reversed, but the normal bundle orientation unchanged.202020To see that the normal orientation has to be the same in 𝒟\mathcal{D} and 𝒟¯\overline{\mathcal{D}}, consider the statement of reflection positivity when 𝒟\mathcal{D} is normal to the t=0t=0 time slice. We will assume that this axiom is satisfied.

The Hermiticity axiom leads to reality conditions on the couplings λ𝒪\lambda_{\mathcal{O}} in the effective action SeffS_{\text{eff}}. In the simplest scenario, whatever parities are broken by 𝒟Σ\mathcal{D}_{\Sigma}, the respective orientations are determined in some way from the orientations of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. This means that if

𝒟1𝒟2𝒟Σ[eSeff]\displaystyle\mathcal{D}_{1}\mathcal{D}_{2}\approx\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}}] (120)

then

𝒟¯1𝒟¯2𝒟¯Σ[eSeff¯],\displaystyle\overline{\mathcal{D}}_{1}\overline{\mathcal{D}}_{2}\approx\overline{\mathcal{D}}_{\Sigma}[e^{-\overline{S_{\text{eff}}}}], (121)

where Seff¯\overline{S_{\text{eff}}} applies complex conjugation and reverses all orientations212121The bulk and the defect orientations, but not the normal bundle orientation. that were used to construct the action. Using this in the Hermiticity axiom for 𝒵(M,𝒟1,𝒟2)\mathcal{Z}(M,\mathcal{D}_{1},\mathcal{D}_{2}) we deduce that

Seff=Seff¯\displaystyle S_{\text{eff}}=\overline{S_{\text{eff}}} (122)

on any manifold MM. This implies that parity-even terms in SeffS_{\text{eff}} have to be real, while parity-odd terms, if at all allowed, have to be imaginary. Here, parity-even and parity-odd refers to the behaviour under the reversal of bulk and defect orientations, but not the normal bundle orientation.

In principle it might be possible that the orientation of 𝒟Σ\mathcal{D}_{\Sigma} is obtained from some curvature invariants rather than from the orientations of 𝒟i\mathcal{D}_{i}, in which case (121) might require a more careful treatment.222222In other words, the defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} might be fully-parity preserving, but the particular fusion configuration might be chiral. E.g. already the fusion of two unoriented distinct line defects in d=3d=3 breaks the bulk parity by the quantity v[μ()νvλ]v^{[\mu}(\nabla^{\perp})^{\nu}v^{\lambda]}. However, this requires a rather exotic behaviour where the theory 𝒟Σ\mathcal{D}_{\Sigma} is different for flat and curved settings, which can only happen under certain conditions. See section 3.8 for a discussion of this in the context of transverse symmetry breaking.

3.3 Weyl invariance

The action (116) is constrained by diffeomorphism and Weyl invariance. If we ignore the possible anomalies (we return to them in section 6), SeffS_{\text{eff}} has to be Weyl- and diffeomorphism-invariant.

While it is easy to build diffeomorphism invariants by properly contracting indices, the construction of Weyl invariants is, in general, a much more subtle task (see e.g. Fefferman:2007rka ). Fortunately, as we have shown in section 2.3, the geometry of a pair of defects allows us to construct a Weyl-weight 1 scalar function (x)\ell(x) in a neighbourhood of 𝒟1\mathcal{D}_{1}. On 𝒟1\mathcal{D}_{1} this function restricts to a measure of the local distance between 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} and in particular (x)L\ell(x)\sim L. The non-trivial statement of section 2.3 is that (x)\ell(x) is also defined in the bulk close to 𝒟1\mathcal{D}_{1}.

Under a Weyl transformation gμν(x)e2ω(x)gμν(x)g_{\mu\nu}(x)\to e^{2\omega(x)}g_{\mu\nu}(x) this function transforms as

(x)eω(x)(x).\displaystyle\ell(x)\to e^{\omega(x)}\ell(x). (123)

In other words, log(x)\log\ell(x) plays the role of a “dilaton”. It allows us to define a new metric

g^μν(x)=(x)2gμν(x)\displaystyle\widehat{g}_{\mu\nu}(x)=\ell(x)^{-2}g_{\mu\nu}(x) (124)

which is invariant under the Weyl transformations of gg. As explained in section 2.3, any Weyl invariant constructed from gg can be written as a diffeomorphism invariant of g^\widehat{g}. We will use hat to denote various geometric quantities in g^\widehat{g}, e.g. \RomanbarII^abμ\widehat{\Romanbar{II}}^{\mu}_{ab} is the second fundamental form. Note that in this metric vμv^{\mu} has unit length,

g^μνvμvν=1.\displaystyle\widehat{g}_{\mu\nu}v^{\mu}v^{\nu}=1. (125)

We will often raise, lower and contract indices using g^μν\widehat{g}_{\mu\nu} instead of gμνg_{\mu\nu}. It should always be clear from the context which metric is used.

In order for (116) to be Weyl-invariant, the couplings λ𝒪\lambda_{\mathcal{O}} will have to transform non-trivially under Weyl transformations to compensate for the transformations of the operators 𝒪{\mathcal{O}} and the measure γ\sqrt{\gamma}. It is therefore convenient to introduce new Weyl-invariant couplings using the following procedure. Firstly, we recall that there is no need to include descendants in (116). Secondly, for primary operators we define232323The shifts by the number of indices follow from the relation (268) between the scaling dimension and the Weyl weight. The indices on 𝒪{\mathcal{O}} have to be raised and lowered by gμνg_{\mu\nu} while the indices on 𝒪^\widehat{{\mathcal{O}}} are raised and lowered by g^μν\widehat{g}_{\mu\nu}.

𝒪^ν1νnμ1μmΔ𝒪n+m𝒪ν1νnμ1μm,\displaystyle\widehat{{\mathcal{O}}}^{\mu_{1}\cdots\mu_{m}}_{\nu_{1}\cdots\nu_{n}}\equiv\ell^{\Delta_{\mathcal{O}}-n+m}{\mathcal{O}}^{\mu_{1}\cdots\mu_{m}}_{\nu_{1}\cdots\nu_{n}}, (126)

where the indices can be normal or defect indices. The operators 𝒪^\widehat{{\mathcal{O}}} are invariant under Weyl transformations of gμνg_{\mu\nu}.242424Anomalous behaviour under Weyl transformations is possible; this is a scheme-dependent question that we discuss in section 3.7.

The effective action (116) can now be written as (note that 𝟏^=𝟏\widehat{\mathbf{1}}=\mathbf{1})

Seff=dpzγ^(λ1^(z)𝟏+𝒪^λ𝒪^(z)𝒪^(z)).\displaystyle S_{\text{eff}}=\int d^{p}z\sqrt{\widehat{\gamma}}\left(\lambda_{\widehat{1}}(z)\mathbf{1}+\sum_{\widehat{{\mathcal{O}}}}\lambda_{\widehat{{\mathcal{O}}}}(z)\widehat{{\mathcal{O}}}(z)\right). (127)

Weyl-invariance of this action is now equivalent to Weyl-invariance of λ𝒪^\lambda_{\widehat{{\mathcal{O}}}}. Here we keep the indices implicit; if 𝒪^\widehat{{\mathcal{O}}} has spin indices, they have to be contracted with dual indices in λ𝒪^\lambda_{\widehat{{\mathcal{O}}}}. The couplings λ𝒪^\lambda_{\widehat{{\mathcal{O}}}} can be constructed as general diff invariants of g^μν,vμ\widehat{g}_{\mu\nu},v^{\mu}, and the shape of 𝒟1\mathcal{D}_{1}. The constraints on g^\widehat{g} discussed in section 2.4 have to be taken into account.

3.4 Cosmological constant term, defect spectrum, and unitarity

The leading allowed term in the effective action is the no-derivatives part of λ1^\lambda_{\widehat{1}}, i.e. the leading coupling to the identity operator. The only possible term is

Seff(0)=a0dpzγ^,\displaystyle S_{\text{eff}}^{(0)}=-a_{0}\int\,d^{p}z\,\sqrt{\widehat{\gamma}}, (128)

where a0a_{0}\in\mathbb{R} is a Wilson coefficient, and the sign is introduced for future convenience. Here and in what follows, we often do not write out the identity operator 𝟏\mathbf{1} explicitly. In terms of the original metric this becomes

Seff(0)=a0dpzγpLp.\displaystyle S_{\text{eff}}^{(0)}=-a_{0}\int\,d^{p}z\,\sqrt{\gamma}\ell^{-p}\sim L^{-p}. (129)

This term has a simple interpretation in terms of the Hilbert space. For example, assume that p=1p=1 and consider the theory on the cylinder M=×Sd1M=\mathbb{R}\times S^{d-1}, where 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are running along \mathbb{R}, at fixed positions on Sd1S^{d-1}. Let HH be the Hamiltonian generating the time translations along \mathbb{R}. We can then ask how the spectrum of HH behaves as the distance between 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} goes to 0.

Translated to this language, the relation (115) states that the spectrum of HH can be approximated by

Hspectruma01𝟏+O(0)𝟏+HΣ+small,\displaystyle H\overset{\text{spectrum}}{\approx}-a_{0}\ell^{-1}\mathbf{1}+O(\ell^{0})\mathbf{1}+H_{\Sigma}+\text{small}, (130)

where the “small” terms contain irrelevant (or marginal with O(L)O(L) couplings) operators and tend to 0 as L0L\to 0, while HΣH_{\Sigma} is the Hamiltonian for the setup in which 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are replaced by 𝒟Σ\mathcal{D}_{\Sigma}. The right-hand side above acts on the Hilbert space of 𝒟Σ\mathcal{D}_{\Sigma}. This shows that a01-a_{0}\ell^{-1} gives the leading contribution to the ground state energy in the limit where 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} get close to each other. If we treat 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} as representing external particles added into the system, a01-a_{0}\ell^{-1} becomes the leading term in the effective interaction potential between these particles, as mediated by the CFT. This picture obviously generalises to p>1p>1 in which case external point particles are replaced by some extended objects. In p=1p=1 case the ground state energy a01+O(1)-a_{0}\ell^{-1}+O(1) of HH is related to cusp anomalous dimension in the limit of deflection angle ϕπ\phi\to\pi,

Γcusp(ϕ)=a0πϕ+O(1).\displaystyle\Gamma_{\text{cusp}}(\phi)=-\frac{a_{0}}{\pi-\phi}+O(1). (131)

We explore this connection further in section 4.

L2\frac{L}{2}L2\frac{L}{2}x1=0x^{1}=0𝒟2\mathcal{D}_{2}𝒟1\mathcal{D}_{1}
Figure 4: The reflection positive set-up for theorem 3.1.

It is easy to show that in a unitary CFT the effective potential is attractive if 𝒟1𝒟2¯\mathcal{D}_{1}\simeq\overline{\mathcal{D}_{2}} (see section 3.2 for the meaning of the conjugation).

Theorem 3.1.

If 𝒟2\mathcal{D}_{2} is conjugate to 𝒟1\mathcal{D}_{1} then a00a_{0}\geq 0.

Proof.

Consider the configuration where 𝒟1\mathcal{D}_{1} is a pp-sphere of radius 1 lying in the hyperplane x1=L/2x^{1}=-L/2, and 𝒟2\mathcal{D}_{2} is obtained from it by applying the Hermitian conjugation in the quantisation where Euclidean time t=x1t=x^{1}. Then 𝒟2\mathcal{D}_{2} is at x1=+L/2x^{1}=+L/2, see figure 4. Furthermore,

f(L)𝒟2𝒟1=Ψ(L)|Ψ(L)0,\displaystyle f(L)\equiv\langle\mathcal{D}_{2}\mathcal{D}_{1}\rangle=\langle\Psi(L)|\Psi(L)\rangle\geq 0, (132)

where |Ψ(L)|\Psi(L)\rangle is the state on x1=0x^{1}=0 hyperplane created by the path integral over the half-space containing 𝒟1\mathcal{D}_{1}. We claim that f(L)f(L) is a monotonically decreasing function of LL. Indeed,

f(L+δL)=Ψ(L)|eδLH|Ψ(L),\displaystyle f(L+\delta L)=\langle\Psi(L)|e^{-\delta LH}|\Psi(L)\rangle, (133)

where HH is the Hamiltonian appropriate for our quantisation. Since H0H\geq 0, it follows that eδLH1\|e^{-\delta LH}\|\leq 1 for δL>0\delta L>0 and thus

f(L+δL)=Ψ(L)|eδLH|Ψ(L)Ψ(L)|Ψ(L)=f(L).\displaystyle f(L+\delta L)=\langle\Psi(L)|e^{-\delta LH}|\Psi(L)\rangle\leq\langle\Psi(L)|\Psi(L)\rangle=f(L). (134)

On the other hand, for small LL we have

logf(L)=a0volSpLp+\displaystyle\log f(L)=a_{0}\mathop{\mathrm{vol}}S^{p}L^{-p}+\cdots (135)

which can only be decreasing for a00a_{0}\geq 0.

This proof as well as the proof of theorem 3.2 below relies on the assumption that 𝒟Σ0\langle\mathcal{D}_{\Sigma}\rangle\neq 0 for a spherical 𝒟Σ\mathcal{D}_{\Sigma} in flat space. If 𝒟Σ=0\langle\mathcal{D}_{\Sigma}\rangle=0 for a unitary defect, then it is easy to show that all correlation functions of a spherical 𝒟Σ\mathcal{D}_{\Sigma} in flat space with arbitrary insertions on and off 𝒟Σ\mathcal{D}_{\Sigma} have to vanish. ∎

As an example, consider the free Maxwell theory in 4d, with 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} being Wilson lines with charges q1q_{1} and q2q_{2}. Reversing the orientation of a Wilson line is equivalent to changing the sign of the charge, so we partially fix this ambiguity by requiring that the orientations of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are such that they agree in the limit L0L\to 0. Theorem 3.1 then says that if q1=q2q_{1}=-q_{2} then the potential between these Wilson lines is attractive. In fact, we have in this case

a0=q1q2=q12>0.\displaystyle a_{0}=-q_{1}q_{2}=q_{1}^{2}>0. (136)

Let us denote the defect conjugate to 𝒟i\mathcal{D}_{i} by 𝒟i¯=𝒟i¯\mathcal{D}_{\overline{i}}=\overline{\mathcal{D}_{i}}, and the constant a0a_{0} appearing in the fusion of 𝒟i\mathcal{D}_{i} with 𝒟j\mathcal{D}_{j} by a0ija_{0}^{ij}. In the Maxwell theory example we then have qi¯=qiq_{\overline{i}}=-q_{i} and a0ij=qiqja_{0}^{ij}=-q_{i}q_{j}. In this case it follows from the arithmetic-geometric mean inequality that

a0ij=qiqjqi2+qj22=a0ii¯+a0jj¯2.\displaystyle a_{0}^{ij}=-q_{i}q_{j}\leq\frac{q_{i}^{2}+q_{j}^{2}}{2}=\frac{a^{i\overline{i}}_{0}+a^{j\overline{j}}_{0}}{2}. (137)

In other words, the charges ii and jj cannot attract more strongly than the average attraction in pairs ii¯i\overline{i} and jj¯j\overline{j}. This result holds more generally:

Theorem 3.2.

For any pair of defects 𝒟i\mathcal{D}_{i} and 𝒟j\mathcal{D}_{j} the following inequality holds,

a0ija0ii¯+a0jj¯2.\displaystyle a^{ij}_{0}\leq\frac{a^{i\overline{i}}_{0}+a^{j\overline{j}}_{0}}{2}. (138)

Note that the right-hand side is always non-negative by theorem 3.1.

Proof.

This follows similarly to theorem 3.1 from the L0L\to 0 limit of the Cauchy-Schwarz inequality

|Ψi¯(L)|Ψj(L)|2|Ψi¯(L)|Ψi¯(L)||Ψj(L)|Ψj(L)|,\displaystyle|\langle\Psi_{\overline{i}}(L)|\Psi_{j}(L)\rangle|^{2}\leq|\langle\Psi_{\overline{i}}(L)|\Psi_{\overline{i}}(L)\rangle||\langle\Psi_{j}(L)|\Psi_{j}(L)\rangle|, (139)

where |Ψi(L)|\Psi_{i}(L)\rangle is created by 𝒟i\mathcal{D}_{i}. ∎

A common scenario is when 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} (or 𝒟¯2\overline{\mathcal{D}}_{2}) belong to the same sufficiently symmetric conformal manifold. In this case a0a_{0} becomes a function a0(s)a_{0}(s) of the distance ss between 𝒟1\mathcal{D}_{1} and 𝒟2{\mathcal{D}_{2}} on this conformal manifold. This is the case for magnetic line defects in O(N)O(N) models and for Maldacena-Wilson lines in 𝒩=4\mathcal{N}=4 SYM, where in both cases ss becomes the angle between the scalar couplings. Theorem 3.2 then implies that a0(0)a_{0}(0) should be the absolute maximum of the function a0(s)a_{0}(s). This is indeed satisfied by explicit results for the Wilson lines at weak and strong coupling Maldacena:1998im ; Drukker:1999zq , and for O(N)O(N) models in ϵ\epsilon-expansion Diatlyk:2024zkk .

3.5 𝒟Σ\mathcal{D}_{\Sigma} is generically simple

One objection that can be raised against (115) is that the right-hand side involves a single defect 𝒟Σ\mathcal{D}_{\Sigma} rather than a sum over contributions from many defects. After all, this is what happens in the local operator product expansion and also in the fusion of topological defects.

The resolution to this is twofold. Firstly, sums of defects of the form

α1𝒟Σ1+α2𝒟Σ2+\displaystyle\alpha_{1}\mathcal{D}_{\Sigma_{1}}+\alpha_{2}\mathcal{D}_{\Sigma_{2}}+\cdots (140)

with arbitrary coefficients αi\alpha_{i} are not possible, since considering α1𝒟Σ1\alpha_{1}\mathcal{D}_{\Sigma_{1}} amounts to adding logα1-\log\alpha_{1} to the action of 𝒟S1\mathcal{D}_{S_{1}}. But logα1-\log\alpha_{1} is not an integral of a local quantity over the defect and thus will break locality of the defect in the sense of the usual cutting and gluing rules for the path integral. It is however possible to consider direct sums of defects

𝒟Σ1+𝒟Σ2+.\displaystyle\mathcal{D}_{\Sigma_{1}}+\mathcal{D}_{\Sigma_{2}}+\cdots. (141)

This can formally be expressed by saying that we allow 𝒟Σ\mathcal{D}_{\Sigma} to be non-simple. Fusion of topological defects generally leads non-simple defects which decompose into direct sums of simple defects.

Secondly, allowing non-simple 𝒟Σ\mathcal{D}_{\Sigma} generically requires going beyond perturbative effective field theory description in the sense that we describe below. If we assume

𝒟Σ=i𝒟Σi,\displaystyle\mathcal{D}_{\Sigma}=\sum_{i}\mathcal{D}_{\Sigma_{i}}, (142)

where 𝒟Σi\mathcal{D}_{\Sigma_{i}} are simple, then the effective action (127) can include local operators from any of the 𝒟Σi\mathcal{D}_{\Sigma_{i}}, including a separate identity operator for each of them.252525There is a subtlety in the case of line defects which we discuss below. In particular, this means that each 𝒟Σi\mathcal{D}_{\Sigma_{i}} comes with its own cosmological constant term a0,ia_{0,i}, and we have no general reason to expect any coincidences between the values of the a0,ia_{0,i}. In other words, generically we expect that a0,ia_{0,i} are all distinct numbers. As will become clear momentarily, even if the number of defects is infinite, we still expect a0,ia_{0,i} to be bounded from above. Without loss of generality, we can then assume that the defects are labelled so that a0,1>a0,2>a_{0,1}>a_{0,2}>\cdots.

In this situation, the defect 𝒟Σ1\mathcal{D}_{\Sigma_{1}} will give the leading contribution to any partition function, and the defects 𝒟Σi\mathcal{D}_{\Sigma_{i}} with i>0i>0 will be suppressed by exponential amounts ec(a0,1a0,i)Lpe^{-c(a_{0,1}-a_{0,i})L^{-p}} for some c>0c>0. Note that existence of an upper bound on a0,ia_{0,i} is therefore required to have a meaningful leading term. The conformal perturbation theory on 𝒟Σ1\mathcal{D}_{\Sigma_{1}} is at any order only power-law suppressed in LL relative to the leading term in 𝒟Σ1\mathcal{D}_{\Sigma_{1}}, and therefore the contributions of 𝒟Σi\mathcal{D}_{\Sigma_{i}} with i2i\geq 2 have to be viewed as non-perturbative effects.

We conclude that if we stay within the framework of perturbative effective field theory, we can generically assume that 𝒟Σ\mathcal{D}_{\Sigma} is a simple defect. This can only be violated if there is some mechanism which makes the leading couplings262626The nn-derivative couplings to identity operator with n<pn<p, scaling as negative powers of LL. coincide for two defects. The latter happens, for example, for topological defects, where their topological nature sets all these couplings to zero.

Another example is the fusion of conformal defects in 2d Ising model, studied in Bachas:2013ora . These conformal defects are denoted 𝒟𝒪,Λ\mathcal{D}_{{\mathcal{O}},\Lambda}, labelled by a Virasoro primary 𝒪{1,ϵ,σ}{\mathcal{O}}\in\{1,\epsilon,\sigma\} and a matrix ΛO(1,1)/2\Lambda\in\mathrm{O}(1,1)/\mathbb{Z}_{2}. For Λ=1\Lambda=1 the defects are topological, with 𝒟1,1\mathcal{D}_{1,1} being the trivial defect, 𝒟ϵ,1\mathcal{D}_{\epsilon,1} the 2\mathbb{Z}_{2} symmetry defect, and 𝒟σ,1\mathcal{D}_{\sigma,1} the Kramers-Wannier duality defect. For Λ1\Lambda\neq 1, the conformal defects can be viewed as deformations of these topological defects. In Bachas:2013ora it was shown that the fusion rules for the defects are

𝒟𝒪1,Λ1×𝒟𝒪2,Λ2𝒟𝒪1×𝒪2,Λ1Λ2,\displaystyle\mathcal{D}_{{\mathcal{O}}_{1},\Lambda_{1}}\times\mathcal{D}_{{\mathcal{O}}_{2},\Lambda_{2}}\to\mathcal{D}_{{\mathcal{O}}_{1}\times{\mathcal{O}}_{2},\Lambda_{1}\Lambda_{2}}, (143)

where 1×𝒪=𝒪,σ×ϵ=σ,ϵ×ϵ=1,σ×σ=1+ϵ1\times{\mathcal{O}}={\mathcal{O}},\sigma\times\epsilon=\sigma,\epsilon\times\epsilon=1,\sigma\times\sigma=1+\epsilon and the right-hand side is understood as the IR defect 𝒟Σ\mathcal{D}_{\Sigma} that is obtained from the fusion in the left-hand side. In particular,

𝒟σ,Λ1×𝒟σ,Λ2𝒟1,Λ1Λ2+𝒟ϵ,Λ1Λ2.\displaystyle\mathcal{D}_{\sigma,\Lambda_{1}}\times\mathcal{D}_{\sigma,\Lambda_{2}}\to\mathcal{D}_{1,\Lambda_{1}\Lambda_{2}}+\mathcal{D}_{\epsilon,\Lambda_{1}\Lambda_{2}}. (144)

Therefore, in this case the fusion of conformal defects yields a non-simple defect. The equality between the coefficients a0a_{0} for 𝒟1,Λ1Λ2\mathcal{D}_{1,\Lambda_{1}\Lambda_{2}} and 𝒟ϵ,Λ1Λ2\mathcal{D}_{\epsilon,\Lambda_{1}\Lambda_{2}} can be explained by considering the fusion with the topological 2\mathbb{Z}_{2} symmetry defect 𝒟ϵ,1\mathcal{D}_{\epsilon,1}. It leaves the left-hand side in (144) invariant, while exchanging the defects 𝒟1,Λ1Λ2\mathcal{D}_{1,\Lambda_{1}\Lambda_{2}} and 𝒟ϵ,Λ1Λ2\mathcal{D}_{\epsilon,\Lambda_{1}\Lambda_{2}} in the right-hand side.

An interesting subtlety appears in the case of line defects. The local operators on a direct sum defect also include off-diagonal operators that come from “defect-changing” local operators that connect 𝒟Σi\mathcal{D}_{\Sigma_{i}} with 𝒟Σj\mathcal{D}_{\Sigma_{j}} (see Nagar:2024mjz for a recent explicit appearance). For concreteness, consider the case when

𝒟Σ=𝒟Σ1+𝒟Σ2,\displaystyle\mathcal{D}_{\Sigma}=\mathcal{D}_{\Sigma_{1}}+\mathcal{D}_{\Sigma_{2}}, (145)

and let δa0=a0,1a0,20\delta a_{0}=a_{0,1}-a_{0,2}\geq 0. Suppose that we have an operator 𝒪12{\mathcal{O}}_{12} that appears in the junction 𝒟Σ1𝒟Σ2\mathcal{D}_{\Sigma_{1}}\mathcal{D}_{\Sigma_{2}}. Its conjugate 𝒪¯12\overline{{\mathcal{O}}}_{12} then appears in the junction 𝒟Σ2𝒟Σ1\mathcal{D}_{\Sigma_{2}}\mathcal{D}_{\Sigma_{1}}. In conformal perturbation theory we will get integrated correlators of products of 𝒪12{\mathcal{O}}_{12} and 𝒪¯12\overline{{\mathcal{O}}}_{12} inserted on 𝒟Σ\mathcal{D}_{\Sigma} in an alternating pattern. In between 𝒪¯12\overline{{\mathcal{O}}}_{12} and 𝒪12{\mathcal{O}}_{12} we will have fragments of 𝒟Σ1\mathcal{D}_{\Sigma_{1}} and in between 𝒪12{\mathcal{O}}_{12} and 𝒪¯12\overline{{\mathcal{O}}}_{12} we will have fragments of 𝒟Σ2\mathcal{D}_{\Sigma_{2}}.

A fragment of 𝒟Σ2\mathcal{D}_{\Sigma_{2}} of length ss will add relative suppression by a factor eδa0s/Le^{-\delta a_{0}s/L}, and thus the typical contributions will have fragments of 𝒟2\mathcal{D}_{2} of length L/δa0L/\delta a_{0}. In a generic situation δa0=O(1)\delta a_{0}=O(1) and this is a UV effect. The short fragments of 𝒟Σ2\mathcal{D}_{\Sigma_{2}} can be replaced using OPE by local operators on 𝒟Σ1\mathcal{D}_{\Sigma_{1}}, and the effective theory at the IR scale RR can be expressed solely in terms of operators on 𝒟Σ1\mathcal{D}_{\Sigma_{1}}.

Suppose now that there is a tunable parameter λ\lambda and δa0=c0λ+O(λ2)\delta a_{0}=c_{0}\lambda+O(\lambda^{2}) with c=O(1)c=O(1). Then the typical size of 𝒟Σ2\mathcal{D}_{\Sigma_{2}} fragments is L/λL/\lambda and for λ1\lambda\ll 1 the effective theory in terms of 𝒟Σ1\mathcal{D}_{\Sigma_{1}} is valid only at scales RL/λR\gg L/\lambda. On the other hand, the effective theory in terms of 𝒟Σ1+𝒟Σ2\mathcal{D}_{\Sigma_{1}}+\mathcal{D}_{\Sigma_{2}} is valid for RLR\gg L. In connection to the cusp anomalous dimension in ×Sd1\mathbb{R}\times S^{d-1} picture (see section 4.1), L=αL=\alpha and R=1R=1 is the radius of the spatial sphere. Thus for λ1\lambda\ll 1 the “simple” 𝒟Σ1\mathcal{D}_{\Sigma_{1}} effective theory should be used to describe αλ\alpha\ll\lambda, while if α1\alpha\ll 1 only then 𝒟Σ1+𝒟Σ2\mathcal{D}_{\Sigma_{1}}+\mathcal{D}_{\Sigma_{2}} should be used.

This discussion applies, for example, to Wilson lines in gauge theories. In this case λ\lambda is the gauge coupling, 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are the fundamental and anti-fundamental lines, while 𝒟Σ1\mathcal{D}_{\Sigma_{1}} and 𝒟Σ2\mathcal{D}_{\Sigma_{2}} are the trivial line and the adjoint line, see e.g. Pineda:2007kz ; Correa:2012nk in the context of 𝒩=4\mathcal{N}=4 SYM.272727Confusingly from our point of view, these references call LL the soft scale, and L/λL/\lambda the ultra-soft. The discussion above appears to be a more abstract version of the pNRQCD effective theory discussed in Pineda:2007kz , although we have not tried to make this connection precise.

If 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are higher-dimensional defects, the contribution of 𝒟2\mathcal{D}_{2} is suppressed by eδa0Rp/Lpe^{-\delta a_{0}R^{p}/L^{p}}, and so if δa0=c0λ+O(λ)\delta a_{0}=c_{0}\lambda+O(\lambda) we cannot ignore 𝒟2\mathcal{D}_{2} unless RL/λ1/pR\gg L/\lambda^{1/p}. However, in this case there is no non-trivial mixing of the two defects.

An important caveat to the above discussion is that we do not know if keeping 𝒟2\mathcal{D}_{2} makes sense for δa00\delta a_{0}\neq 0. In general there could be other non-perturbative in LL effects that are equally or more important, effectively destroying 𝒟2\mathcal{D}_{2} as a well-defined defect.

3.6 Subleading terms in the identity operator coupling

We now consider the subleading terms in the coupling λ1^\lambda_{\widehat{1}} of the identity operator. In general, their form depends on the dimension pp of the defects and on the dimension dd of the bulk CFT. Here we will consider only the one-derivative terms. The two-derivative terms for line defects and for codimension q=1q=1 and q=2q=2 defects are classified in appendix C.

First of all, it is easy to see that for p>1p>1 no one-derivative terms exist. Indeed, the only geometric one-derivative tensor at our disposal is \RomanbarII^abμ\widehat{\Romanbar{II}}^{\mu}_{ab} and we can also form the covariant derivative ^avμ\widehat{\nabla}_{a}^{\perp}v^{\mu}.282828Note that we do not count the derivative in aX1μ\partial_{a}X_{1}^{\mu}. This is because X1μX_{1}^{\mu} can only appear in this combination, and X1X_{1} itself carries a factor of the IR length scale RR. However, there is nothing to contract their defect indices with — recall that \RomanbarII^abμ\widehat{\Romanbar{II}}^{\mu}_{ab} is traceless as required by (82).

On line defects, however, it is possible to introduce the unit tangent vector field tat^{a} on 𝒟1\mathcal{D}_{1}. It defines an orientation of 𝒟Σ\mathcal{D}_{\Sigma}; if enough parity is preserved then the effective action has to be an even function of tt. If d>3d>3, this vector doesn’t change the situation since \RomanbarII^abμtatb=\RomanbarII^μ=0\widehat{\Romanbar{II}}^{\mu}_{ab}t^{a}t^{b}=\widehat{\Romanbar{II}}^{\mu}=0 and v^tv=0v\cdot\widehat{\nabla}^{\bot}_{t}v=0. Therefore,

Seff=dpzγ^(a0+2-derivative terms)+(p>1p=1,d>3),\displaystyle S_{\text{eff}}=\int\,d^{p}z\,\sqrt{\widehat{\gamma}}\left(-a_{0}+\text{2-derivative terms}\right)+\cdots\quad\left(p>1\atop p=1,d>3\right), (146)

where \cdots represents non-identity operators. For p=1p=1 all the omitted terms in (146) vanish in the limit L0L\to 0.

If d=3d=3 then we can define a unit vector uμu^{\mu} that is orthogonal to both vμv^{\mu} and tμt^{\mu}, i.e. uμ=ε^μvμρσtσu^{\mu}=\widehat{\varepsilon}^{\mu}{}_{\rho\sigma}v^{\mu}t^{\sigma}. This allows us to construct a non-trivial term at one-derivative order,

Seff=𝑑zγ^(a0+ia1u^tv+2-derivative terms)+(p=1,d=3),\displaystyle S_{\text{eff}}=\int\,dz\,\sqrt{\widehat{\gamma}}\left(-a_{0}+ia_{1}u\cdot\widehat{\nabla}^{\perp}_{t}v+\text{2-derivative terms}\right)+\cdots\quad(p=1,\,d=3), (147)

where a1a_{1}\in\mathbb{R} (see sec 3.2). This term breaks parity and requires either the bulk orientation or both the defect and the normal bundle orientations.

Example: pure Chern-Simons theory

The term multiplying 2πia12\pi i\,a_{1},

I(v)\displaystyle I(v) =12π𝑑zγ^u^tv=12π𝑑zγ^ϵ^μνλvνtλ^tvμ\displaystyle=\frac{1}{2\pi}\int\,dz\,\sqrt{\widehat{\gamma}}u\cdot\widehat{\nabla}^{\perp}_{t}v=\frac{1}{2\pi}\int\,dz\,\sqrt{\widehat{\gamma}}\widehat{\epsilon}_{\mu\nu\lambda}v^{\nu}t^{\lambda}\widehat{\nabla}^{\perp}_{t}v^{\mu}
=12π𝑑zγϵμνλvντλτvμ,\displaystyle=\frac{1}{2\pi}\int\,dz\,\sqrt{\gamma}\epsilon_{\mu\nu\lambda}v^{\nu}\tau^{\lambda}\nabla^{\perp}_{\tau}v^{\mu}, (148)

where τ\tau is a tangent unit vector for the physical metric, contains information about the topological linking of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}.292929Note that I(v)I(v) is defined on any orientable manifold (provided 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are close enough to each other), while the linking number is only defined if 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are null homologous. However, I(v)I(v) itself isn’t topologically-invariant. A useful description of it is the following. Fix a normal vector field v0v_{0} and the corresponding u0μ=ϵ^μv0ννλtλu_{0}^{\mu}=\widehat{\epsilon}^{\mu}{}_{\nu\lambda}v_{0}^{\nu}t^{\lambda}. Any unit normal vector field vv can be written as

v=cosθv0+sinθu0,\displaystyle v=\cos\theta v_{0}+\sin\theta u_{0}, (149)

for some function θ(z)\theta(z). In terms of θ\theta we have

I(v)=12π𝑑zdθdz+I(v0)+I(v0).\displaystyle I(v)=\frac{1}{2\pi}\int dz\frac{d\theta}{dz}+I(v_{0})\in\mathbb{Z}+I(v_{0}). (150)

Therefore, I(v)I(v) is a homotopy invariant of vv and I(v)I(v0)I(v)-I(v_{0}) is an integer which describes the difference between the linking numbers of vv and v0v_{0} with 𝒟1\mathcal{D}_{1}. However, I(v)mod1I(v)\!\mod 1 is vv-independent and, as can be seen by setting v=\RomanbarII/|\RomanbarII|v=\Romanbar{II}/|\Romanbar{II}|, is proportional to the total torsion of the curve in the physical metric, which is not a topological invariant and can be equal to any real number even in flat space.

Despite I(v)I(v) not being exactly topological, it is relevant for fusion of Wilson lines in 3d pure Chern-Simons theory, which we now consider. Let 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} be Wilson loops with framing represented by normal vector fields f1f_{1} and f2f_{2}. To keep track of the framing, we will write 𝒟i[fi]\mathcal{D}_{i}[f_{i}]. Recall that the framing anomaly implies that 𝒟i[fi]\mathcal{D}_{i}[f_{i}] is a homotopy invariant of fif_{i}, and whenever f~i\widetilde{f}_{i} differs from fif_{i} by nn 2π2\pi-twists,

𝒟i[f~i]=𝒟i[fi]e2πihin,\displaystyle\mathcal{D}_{i}[\widetilde{f}_{i}]=\mathcal{D}_{i}[f_{i}]e^{2\pi ih_{i}n}, (151)

where hih_{i} is the conformal weight of the corresponding 2d primary Witten:1988hf .

Due to the topological nature of Chern-Simons theory, the only coupling that has a chance of surviving in SeffS_{\text{eff}} is a1a_{1} (0- or higher-derivative couplings explicitly depends on the scale LL). However, two modifications to SeffS_{\text{eff}} are in order. Firstly, we now can use fif_{i} in addition to vv. Secondly, since a0a_{0} vanishes, the defect 𝒟Σ\mathcal{D}_{\Sigma} may not be simple and is in general a direct sum of Wilson lines 𝒟Σk\mathcal{D}_{\Sigma_{k}} (see section 3.5). Correspondingly, there is a separate identity operator for each 𝒟Σk\mathcal{D}_{\Sigma_{k}}, so that the effective action takes the form

Seff=k(2πia1,kI(v)+ib1,kI(f1)+ib2,kI(f2)+ibΣ,kI(fΣ,k))𝟏k,\displaystyle S_{\text{eff}}=\sum_{k}\left(2\pi ia_{1,k}I(v)+ib_{1,k}I(f_{1})+ib_{2,k}I(f_{2})+ib_{\Sigma,k}I(f_{\Sigma,k})\right)\mathbf{1}_{k}, (152)

where fΣ,kf_{\Sigma,k} is the framing vector field for 𝒟Σk\mathcal{D}_{\Sigma_{k}} and a,ba,b are undetermined coefficients.

Matching of the framing anomaly requires setting b1,k=2πh1b_{1,k}=2\pi h_{1}, b2,k=2πh2b_{2,k}=2\pi h_{2} and bΣ,k=2πhΣ,kb_{\Sigma,k}=-2\pi h_{\Sigma,k}. Indeed,

e2πih1I(f1)+2πih2I(f2)2πihΣ,kI(fΣ,k)𝒟Σ,k[fΣ,k]\displaystyle e^{2\pi ih_{1}I(f_{1})+2\pi ih_{2}I(f_{2})-2\pi ih_{\Sigma,k}I(f_{\Sigma,k})}\mathcal{D}_{\Sigma,k}[f_{\Sigma,k}] (153)

reproduces the behaviour of 𝒟1[f1]𝒟2[f2]\mathcal{D}_{1}[f_{1}]\mathcal{D}_{2}[f_{2}] under the change of framing: it transforms as required under changes of f1f_{1} and f2f_{2} and doesn’t depend on fΣ,kf_{\Sigma,k}. While matching the framing anomaly, these couplings introduce dependence on the metric through I()I(\cdot).303030Essentially the same term cancels the framing anomaly in the non-topological regularisation of Polyakov:1988md . This dependence can only be cancelled by the I(v)I(v) term, which requires

a1,k=hΣ,kh1h2.\displaystyle a_{1,k}=h_{\Sigma,k}-h_{1}-h_{2}. (154)

Therefore, the effective action for fusion of Wilson lines is

Seff\displaystyle S_{\text{eff}} =2πik(h1(I(f1)I(v))+h2(I(f2)I(v))hΣ,k(I(fΣ,k)I(v)))𝟏k,\displaystyle=2\pi i\sum_{k}\left(h_{1}(I(f_{1})-I(v))+h_{2}(I(f_{2})-I(v))-h_{\Sigma,k}(I(f_{\Sigma,k})-I(v))\right)\mathbf{1}_{k}, (155)

which is indeed topological. The vv-dependence predicted by this action is easy to derive directly from Chern-Simons theory by considering the state created by 𝒟1[f1]𝒟2[f2]\mathcal{D}_{1}[f_{1}]\mathcal{D}_{2}[f_{2}] on the surface of a thin solid cylinder that contains them, and by studying how this picture transforms under Dehn twists.

3.7 Displacement operator and other contributions

So far we have focused mostly on the couplings to the identity operator. There is nothing qualitatively new when we consider couplings to other operators, except for two comments. Firstly, the contribution of any other operator to the effective action vanishes in the limit L0L\to 0. This is immediate in the case of the irrelevant operators, and for marginal operators this can be achieved by adjusting the definition of 𝒟Σ\mathcal{D}_{\Sigma} as described in section 3.1.

Secondly, the couplings to non-trivial operators enter non-trivially into the conformal perturbation theory. The conformal perturbation theory generally requires renormalisation, and Weyl-invariance can be affected by the choice of the renormalisation scheme. Specifically, the behaviour of the couplings λ𝒪\lambda_{{\mathcal{O}}} in (116) under Weyl transformations may be anomalous. A standard example of this is the Weyl anomaly which can be viewed as an anomalous transformation law for the coupling λ1\lambda_{1} to the identity operator.

If present, anomalous terms in these transformation laws must be polynomial in the derivatives of couplings and respect dimensional analysis. For example, suppose the chosen renormalisation scheme is such that under an infinitesimal Weyl transformation gμν(1+2ω+O(ω2))gμνg_{\mu\nu}\to(1+2\omega+O(\omega^{2}))g_{\mu\nu}, the coupling λ𝒪1\lambda_{{\mathcal{O}}_{1}} has to transform as, schematically,

δωλ𝒪1=Nωλ𝒪1+αmωiλ𝒪ini\displaystyle\delta_{\omega}\lambda_{{\mathcal{O}}_{1}}=N\omega\lambda_{{\mathcal{O}}_{1}}+\alpha\partial^{m}\omega\prod_{i}\lambda_{{\mathcal{O}}_{i}}^{n_{i}} (156)

in order for the partition function to remain Weyl-invariant. Here NN is the Weyl weight of the coupling, ni,m0n_{i},m\geq 0 are integers313131It must be that ini>1\sum_{i}n_{i}>1 since such anomalies come from divergences which cannot arise at linear level in the couplings., α0\alpha\neq 0 is a dimensionless coefficient, and 𝒪i{\mathcal{O}}_{i} are some primary operators, possibly including 𝒪1{\mathcal{O}}_{1}. The derivatives \partial can be actual derivatives acting on the various 𝒪i{\mathcal{O}}_{i} and ω\omega, or instead \partial can represent various tensors such as the curvature tensor, containing an equivalent number of derivatives. Dimensional analysis then requires

pΔ1=m+ini(pΔi),\displaystyle p-\Delta_{1}=m+\sum_{i}n_{i}(p-\Delta_{i}), (157)

where Δi\Delta_{i} is the scaling dimension of 𝒪i{\mathcal{O}}_{i}. Note that this is a rational relation between the scaling dimensions of primary operators 𝒪1{\mathcal{O}}_{1} and 𝒪i{\mathcal{O}}_{i}.

Many such relations exist in free theory, which may introduce subtleties when applying this formalism to weakly-coupled theories. For example, a marginally irrelevant operator would come with a coupling that transforms anomalously according to a corresponding beta function. In generic interacting theories, however, we only expect such relations when protected operators are present. In the rest of this section, we will briefly discuss one generic class of such anomalous terms, associated with the displacement operator DμD_{\mu}.

The displacement operator DμD_{\mu} exists on any local defect, is valued in the normal bundle, and has a protected scaling dimension Δ=p+1\Delta=p+1. The above relations allow anomalous terms of the form δλ𝒪nλDnλ𝒪\delta\lambda_{{\mathcal{O}}}\ni\partial^{n}\lambda_{D}^{n}\lambda_{{\mathcal{O}}}. In particular, the coupling λD\lambda_{D} itself could have an anomalous transformation law.

Therefore, the problem of writing down a Weyl-invariant effective action seems to generically require a determination of such anomalous terms arising from λD\lambda_{D}. Fortunately, we have already implicitly solved this problem in section 2.1. Indeed, a choice of the renormalisation scheme is nothing more than a definition of the partition function as a function of its couplings at non-linear level. In section 2.1 we showed that there exists a Weyl-invariant definition of the displacement coupling λD\lambda_{D}, to all orders in λD\lambda_{D}, and therefore there are no anomalous terms proportional to λD\lambda_{D} in the corresponding renormalisation scheme. While it might not be obvious how to implement this scheme in practice, our arguments do show that such Weyl-invariant schemes exist in principle, to the extent to which Weyl-covariant partition functions can be defined for an arbitrary shape and position of 𝒟Σ\mathcal{D}_{\Sigma}.

Note that above “Weyl-invariant” refers only to the transformation law for λD\lambda_{D} and to anomalous terms of the form δλ𝒪nλDnλ𝒪\delta\lambda_{{\mathcal{O}}}\ni\partial^{n}\lambda_{D}^{n}\lambda_{{\mathcal{O}}}. In general, terms of the form δλ1n+pλDn\delta\lambda_{1}\ni\partial^{n+p}\lambda_{D}^{n} responsible for defect Weyl anomalies are still allowed.

Displacement coupling and the location of 𝒟Σ\mathcal{D}_{\Sigma}

We define the displacement operator on a defect 𝒟\mathcal{D} by the requirement that

𝒟[edpzγvμDμ]\displaystyle\mathcal{D}[e^{-\int d^{p}z\sqrt{\gamma}v^{\mu}D_{\mu}}] (158)

is equivalent to 𝒟\mathcal{D} displaced by the vector field vμv^{\mu} as described in section 2.1. Displacement operator is a local operator of dimension Δ=p+1\Delta=p+1 that can appear in SeffS_{\text{eff}} as any other irrelevant operator.

There is a trade-off between the choice of the insertion point for 𝒟Σ\mathcal{D}_{\Sigma} and the coupling to the displacement operator in SeffS_{\text{eff}}. Indeed, by definition, adding a displacement coupling

Seffdpzγ^λD^μD^μ\displaystyle S_{\text{eff}}\ni\int d^{p}z\sqrt{\widehat{\gamma}}\lambda^{\mu}_{\widehat{D}}\widehat{D}_{\mu} (159)

is completely equivalent to deforming 𝒟Σ\mathcal{D}_{\Sigma} by the vector field λD^μ\lambda^{\mu}_{\widehat{D}}. This means that we are free to choose any configuration for 𝒟Σ\mathcal{D}_{\Sigma} as long as it tends to the common limit of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} as L0L\to 0. Differences between possible choices can be compensated by the coupling λDμ\lambda^{\mu}_{D}. In this paper, we use the convention where the position of 𝒟Σ\mathcal{D}_{\Sigma} coincides with that of 𝒟1\mathcal{D}_{1}. Although not particularly symmetric, this choice simplifies the form of the effective couplings.323232This is similar to the choice one makes when writing down the OPE of two local operators. It is common to choose 𝒪1(x)𝒪2(y)k𝒪k(y){\mathcal{O}}_{1}(x){\mathcal{O}}_{2}(y)\sim\sum_{k}{\mathcal{O}}_{k}(y), but in principle other choices for the coordinates of 𝒪k{\mathcal{O}}_{k} in the right-hand side are possible.

The displacement coupling has the derivative expansion

λD^μ=aD,0vμ+aD,1^tvμ+,\displaystyle\lambda_{\widehat{D}}^{\mu}=a_{D,0}v^{\mu}+a_{D,1}\widehat{\nabla}_{t}^{\perp}v^{\mu}+\cdots, (160)

where the 11-derivative term is only possible for p=1p=1 when one can define the unit tangent vector tμt^{\mu}. The higher-derivative terms contribute at O(L3)O(L^{3}) to the action.

Whenever there is a permutation symmetry between 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}, it is possible to argue that aD,0=12a_{D,0}=\tfrac{1}{2}. In other words, the natural position for 𝒟Σ\mathcal{D}_{\Sigma} is halfway between 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}.

3.8 Breaking of transverse rotations and relevant couplings

In order for the effective action (116) to describe the tail of an RG flow that terminates in 𝒟Σ\mathcal{D}_{\Sigma}, it is necessary that it does not contain any relevant operators. Since in a general defect fusion configuration we do not perform any fine-tuning, this normally requires that 𝒟Σ\mathcal{D}_{\Sigma} must not have any relevant operators that can appear in (116).

When discussing relevant operators, one usually focuses on scalars. However, in our case, due to the presence of vμv^{\mu} and curvature invariants, it is necessary to also consider the defect operators that are charged under transverse rotations. For example, we can have couplings such as

vμ𝒪^1,μ,\RomanbarII^abμ\RomanbarII^νab𝒪^2,μν,\displaystyle v^{\mu}\widehat{{\mathcal{O}}}_{1,\mu},\quad\widehat{\Romanbar{II}}^{\mu}_{ab}\widehat{\Romanbar{II}}^{\nu ab}\widehat{{\mathcal{O}}}_{2,\mu\nu}, (161)

where all the indices are transverse. Taking into account the derivatives in the curvature invariants, the coupling with 𝒪1{\mathcal{O}}_{1} is relevant if Δ1<p\Delta_{1}<p, the coupling with 𝒪2{\mathcal{O}}_{2} is relevant if Δ2<p2\Delta_{2}<p-2.

In the case of bulk RG flows, there is usually no need to consider such couplings since they are excluded by unitarity bounds. For example, in 3 dimensions the smallest possible dimension of a tensor operator is 22 while the simplest curvature invariant is the Riemann tensor, which has 2 derivatives.333333Couplings of Ricci scalar to relevant scalar operators still need to be taken into account even in the bulk. In the defect case, the unitarity bounds do not depend on the transverse spin343434For instance, consider nϕ\partial^{n}\phi-type operators on the trivial pp-dimensional defect in dd-dimensional free scalar theory, where all nn derivatives are transverse and p=d2p=d-2. Their dimension is Δ=n+p/2\Delta=n+p/2, which shows that for any fixed transverse spin, pΔ=p/2np-\Delta=p/2-n can be arbitrarily large if pp is large enough. and therefore this logic does not apply.

If p=1p=1, i.e. when we consider line defects, the only relevant couplings of this kind are

vμ𝒪^μ,vμvν𝒪^μν,\displaystyle v^{\mu}\widehat{{\mathcal{O}}}_{\mu},\quad v^{\mu}v^{\nu}\widehat{{\mathcal{O}}}_{\mu\nu},\cdots (162)

This implies that 𝒟Σ\mathcal{D}_{\Sigma} should not have any (otherwise neutral) relevant operators which are transverse traceless-symmetric tensors.

Starting at least with p=3p=3 surface defects, curvature couplings such as \RomanbarII^abμ\RomanbarII^νab𝒪^2,μν\widehat{\Romanbar{II}}^{\mu}_{ab}\widehat{\Romanbar{II}}^{\nu ab}\widehat{{\mathcal{O}}}_{2,\mu\nu} or \RomanbarII^abμ\RomanbarII^νabvρ𝒪3,μνρ\widehat{\Romanbar{II}}^{\mu}_{ab}\widehat{\Romanbar{II}}^{\nu ab}v^{\rho}{\mathcal{O}}_{3,\mu\nu\rho} can appear. In generic configurations, this prohibits 𝒟Σ\mathcal{D}_{\Sigma} from having relevant operators with transverse spin in certain representations. In special configurations, such as for flat defects, the curvature couplings can vanish, which opens the possibility that the fusion defect 𝒟Σ\mathcal{D}_{\Sigma} can be different for different fusion geometries. For example, one can imagine a situation where for flat defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} the IR defect 𝒟Σ\mathcal{D}_{\Sigma} has a sufficiently relevant transverse mixed-symmetry tensor 𝒪3,μνρ{\mathcal{O}}_{3,\mu\nu\rho}, which cannot couple to a constant vμv^{\mu} alone. When we deform the defects away from the flat configuration, the curvature coupling \RomanbarII^abμ\RomanbarII^νabvρ𝒪3,μνρ\widehat{\Romanbar{II}}^{\mu}_{ab}\widehat{\Romanbar{II}}^{\nu ab}v^{\rho}{\mathcal{O}}_{3,\mu\nu\rho} will turn on and trigger an RG flow to a new IR defect. In other words, fine-tuning of the fusion RG flow might in some cases be achieved by considering defects of special shape.

Related to the issue of relevant operators with transverse spin is possibility that 𝒟Σ\mathcal{D}_{\Sigma} itself breaks the transverse rotations SO(q)\mathrm{SO}(q) down to a proper subgroup HH. Indeed, in the picture where we view the product 𝒟1𝒟2\mathcal{D}_{1}\mathcal{D}_{2} as the UV definition of a non-conformal defect that flows to 𝒟Σ\mathcal{D}_{\Sigma} in the IR, we have to contend with the fact that this UV definition preserves fewer symmetries than the individual defects 𝒟1\mathcal{D}_{1} or 𝒟2\mathcal{D}_{2}. As in the discussion above, this is manifested by presence of vμv^{\mu} and of various curvature invariants.

If 𝒟Σ\mathcal{D}_{\Sigma} does break transverse symmetries down to HH, this means that in order to compute partition functions with 𝒟Σ\mathcal{D}_{\Sigma} inserted, one needs to specify an element n(z)SO(q)/Hn(z)\in\mathrm{SO}(q)/H at every point of the defect. In the simplest case, when H=SO(q1)H=\mathrm{SO}(q-1) and q>2q>2, we have SO(q)/H=Sq1\mathrm{SO}(q)/H=S^{q-1} and nn can be viewed as a unit vector (with respect to gμνg_{\mu\nu}). The only possible value for nn in a fusion process is

n=±1v+O(L).\displaystyle n=\pm\ell^{-1}v+O(L). (163)

The choice of sign here should be viewed as a Wilson coefficient that depends on the particular pair of defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. It can be changed by altering the meaning of nn to n-n for 𝒟Σ\mathcal{D}_{\Sigma}, but this will change the sign in all fusions in which 𝒟Σ\mathcal{D}_{\Sigma} appears (so relative signs are meaningful).

For completeness, we give the form of the first subleading terms for p=1p=1 in dimensions d>3d>3. In general,

n\displaystyle n =±emμνrμvν1v=±1(vμ+rμ(rv)vμ+),\displaystyle=\pm e^{m_{\mu\nu}r^{\mu}v^{\nu}}\ell^{-1}v=\pm\ell^{-1}\left(v^{\mu}+r^{\mu}-(r\cdot v)v^{\mu}+\cdots\right), (164)

where (mρσ)μ=νδρμg^σνδσμg^ρν(m_{\rho\sigma})^{\mu}{}_{\nu}=\delta^{\mu}_{\rho}\widehat{g}_{\sigma\nu}-\delta^{\mu}_{\sigma}\widehat{g}_{\rho\nu} is a rotation generator and rμr^{\mu} is a Weyl weight-0 vector in the normal bundle, defined modulo vμv^{\mu}. The leading order derivative expansion for rr is

rμ=ar,1^tvμ+ar,2(^t)2vμ+modvμ,\displaystyle r^{\mu}=a_{r,1}\widehat{\nabla}^{\perp}_{t}v^{\mu}+a_{r,2}(\widehat{\nabla}^{\perp}_{t})^{2}v^{\mu}+\cdots\mod v^{\mu}, (165)

where tat^{a} is a unit (in g^\widehat{g}) vector tangent to the line defect, aligned with the orientation. Note that we can’t use the Schouten tensor (equivalently, the Ricci tensor) due to (84), and Rvμ=0modvμRv^{\mu}=0\mod v^{\mu}. We also used (82).

Recall that there is a tradeoff between the value of nn and the coupling to the tilt operators in SeffS_{\text{eff}}. When writing the above formula, we assume the convention in which the tilt operators do not appear in SeffS_{\text{eff}}. Alternatively, and this is the convention that we adopt more broadly in this paper, we can set n=±1vn=\pm\ell^{-1}v exactly, and compensate for this by adding the coupling rμr^{\mu} to the tilt operator corresponding to broken transverse rotations. A similar tradeoff is present for any exactly marginal deformation.

In the case d=3d=3 and p=1p=1 the complication is that there is not in general a canonical identification between the elements of SO(2)/SO(1)\mathrm{SO}(2)/\mathrm{SO}(1) and the unit normal vectors, even up to a sign.353535If the defect preserves transverse parity, then we have to work with O(2)/O(1)\mathrm{O}(2)/\mathrm{O}(1) and a canonical identification up to a sign is again possible. In practice this means that we can write

n=Rθ(1v),\displaystyle n=R_{\theta}(\ell^{-1}v), (166)

where RθR_{\theta} is a rotation by angle θ\theta in the normal plane, and we have the derivative expansion

θ=aθ,0+,\displaystyle\theta=a_{\theta,0}+\cdots, (167)

where aθ,0a_{\theta,0} is a Wilson coefficient. Similarly to the sign of nn, we can adjust aθ,0a_{\theta,0} by changing the meaning of nn to Rφ(n)R_{\varphi}(n) for 𝒟Σ\mathcal{D}_{\Sigma}, but this will affect all fusions in which 𝒟Σ\mathcal{D}_{\Sigma} appears, so relative values of aθ,0a_{\theta,0} are meaningful.

Breaking to HH other than SO(q1)\mathrm{SO}(q-1) is not possible if we consider flat parallel defects, as this configuration will explicitly preserve SO(q1)\mathrm{SO}(q-1). So, such HH can be only achieved in the situation of the kind mentioned earlier in this section, when the IR defect 𝒟Σ\mathcal{D}_{\Sigma} is different for generic and flat configurations. Note that in this scenario, the dimension pp of the defect plays a role. For example, for p=1p=1 no relevant curvature couplings can turn on as we move from flat to curved defects. Therefore, we expect that HH smaller than SO(q1)\mathrm{SO}(q-1) cannot be realised for line defects. It is an interesting question whether scenarios of this kind can be realised for higher-dimensional defects.

4 Cusps

In this section we first discuss the general properties of a cusp formed by two line defects with a small opening angle α\alpha, and then focus on the specific example of supersymmetric Wilson lines in 𝒩=4\mathcal{N}=4 SYM.

4.1 General structure of scaling dimensions at a cusp

L2\frac{L}{2}L2\frac{L}{2}x1=0x^{1}=0𝒟2\mathcal{D}_{2}𝒟1\mathcal{D}_{1}𝒟2\mathcal{D}_{2}α\alpha𝒟2\mathcal{D}_{2}
Figure 5: A cusped junction of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}

We consider line defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} and assume that they can be joined at a cusp as in figure 5. Using the standard arguments, operator-state correspondence implies that the eigenstates of the dilatation operator in the Hilbert space of states on Sd1S^{d-1} punctured by 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are in one-to-one correspondence with local operators at the cusp. In particular, one expects infinitely-many choices for local operators that can be present at the cusp.363636The use of “local operator” is a bit confusing since it makes it sound as if it were possible to not insert anything at the cusp. Instead of “local operators at the cusp”, it would perhaps be more useful to talk about different cusped junctions between line defects. Nevertheless, we prefer to use the term “local operator”‘.,373737Note that the Hilbert spaces with different angles α\alpha between 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are all isomorphic via a conformal transformation to the Hilbert space with α=π\alpha=\pi. What really depends on α\alpha is the dilatation operator and its spectrum.

One usually distinguishes the local operator with the lowest scaling dimension. This scaling dimension is denoted Γcusp\Gamma_{\text{cusp}} and the corresponding local operator is often described as the “cusp with no insertions”. In 𝒩=4\mathcal{N}=4 SYM literature it is customary to write Γcusp\Gamma_{\text{cusp}} for fundamental Wilson lines as the function of ϕ=πα\phi=\pi-\alpha.

Via the exponential map, the configuration in figure 5 is conformally equivalent to the cylinder ×Sd1\mathbb{R}\times S^{d-1}, where 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are inserted along \mathbb{R} at fixed points on Sd1S^{d-1} separated by an angle α\alpha. The cusp is mapped to the infinite past. Let HH be the Hamiltonian that generates the time translations along \mathbb{R}. This Hamiltonian is simply the dilatation operator in figure 5 and thus its spectrum is given by the scaling dimensions of cusp local operators (we set the radius of Sd1S^{d-1} to 1). In particular, Γcusp\Gamma_{\text{cusp}} is the ground state energy.

In the limit α0\alpha\to 0 the defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} on ×Sd1\mathbb{R}\times S^{d-1} fuse. Let 𝒟Σ\mathcal{D}_{\Sigma} be the corresponding IR defect. Recall from (130) that the spectrum of HH can be approximated as

H=spectrumHΣ+λ𝟏(α)𝟏+𝒪λ𝒪(α)𝒪,\displaystyle H\overset{\text{spectrum}}{=}H_{\Sigma}+\lambda_{\mathbf{1}}(\alpha)\mathbf{1}+\sum_{{\mathcal{O}}}\lambda_{\mathcal{O}}(\alpha){\mathcal{O}}, (168)

where the sum over 𝒪{\mathcal{O}} is over all primary operators on 𝒟Σ\mathcal{D}_{\Sigma} and HΣH_{\Sigma} is the Hamiltonian when the insertions of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are replaced by 𝒟Σ\mathcal{D}_{\Sigma}. Any possible index contractions are implicit.

The coefficients λ𝒪(α)\lambda_{\mathcal{O}}(\alpha) behave as

λ𝒪(α)αΔ𝒪1.\displaystyle\lambda_{\mathcal{O}}(\alpha)\sim\alpha^{\Delta_{\mathcal{O}}-1}. (169)

Note that by the assumption that 𝒟Σ\mathcal{D}_{\Sigma} is the IR fixed point, the sum in (168) includes only irrelevant operators, so their couplings vanish as α0\alpha\to 0.383838We might have exactly marginal operators with Δ𝒪=1\Delta_{\mathcal{O}}=1. For them we use the convention described in section 3, which makes their effective couplings vanish as at least O(α)O(\alpha).

Using (168) we can in principle predict the small-α\alpha behaviour of the ground-state energy Γcusp\Gamma_{\text{cusp}} (or of any other energy level) to any order in α\alpha using Rayleigh–Schrödinger perturbation theory. The simplest contribution comes from the identity operator, which shifts all energies by λ𝟏(α)\lambda_{\mathbf{1}}(\alpha). In particular, the ground state energy Γcusp\Gamma_{\text{cusp}} is given by

Γcusp=E0,Σ+λ𝟏(α)+δΓcusp,\displaystyle\Gamma_{\text{cusp}}=E_{0,\Sigma}+\lambda_{\mathbf{1}}(\alpha)+\delta\Gamma_{\text{cusp}}, (170)

where E0,ΣE_{0,\Sigma} is the ground-state energy of HΣH_{\Sigma} and δΓcusp\delta\Gamma_{\text{cusp}} is the contribution from non-identity 𝒪{\mathcal{O}} in (168). The latter contribution vanishes in the limit α0\alpha\to 0.

The coupling λ𝟏(α)\lambda_{\mathbf{1}}(\alpha) has an expansion in powers of α\alpha. In appendix C.1 we classify 2-derivative contributions to λ𝟏\lambda_{\mathbf{1}}, see (292). In the conformally-flat case of ×S3\mathbb{R}\times S^{3} the result is, up to two derivatives,

λ𝟏=\displaystyle\lambda_{\mathbf{1}}= 𝑑zγ^(a0+a2,1^tv^tv+a2,2R^+ia2,3P^tv+),\displaystyle\int dz\sqrt{\widehat{\gamma}}\left(-a_{0}+a_{2,1}\widehat{\nabla}_{t}^{\perp}v\cdot\widehat{\nabla}_{t}^{\perp}v+a_{2,2}\widehat{R}+ia_{2,3}\widehat{P}_{tv}+\cdots\right), (171)

where P^μν\widehat{P}_{\mu\nu} is the Schouten tensor and tt is a unit (in the fusion metric) vector field on the line defect. The terms multiplying a2,1a_{2,1} and a2,3a_{2,3} vanish. The derivative ^tv\widehat{\nabla}^{\perp}_{t}v vanishes since the separation between the defects is time-independent. The component P^tv\widehat{P}_{tv} of the Schouten tensor vanishes due to the transverse rotational symmetry of the fusion metric. This symmetry will be present whenever the IR defect is a straight line along \mathbb{R}. This is discussed in more detail in section 2.4 and section 2.5 where we also show that for this configuration R^=(d1)α2\widehat{R}=-(d-1)\alpha^{2}, see equation (114).

This shows that

λ𝟏(α)=a0α(d1)a2,2α+.\displaystyle\lambda_{\mathbf{1}}(\alpha)=-\frac{a_{0}}{\alpha}-(d-1)a_{2,2}\alpha+\cdots. (172)

In fact, it is easy to see that λ𝟏(α)\lambda_{\mathbf{1}}(\alpha) has an expansion in odd powers of α\alpha. This is because it is given by α1\alpha^{-1} times local invariants which are all constructed from the derivatives of the fusion metric as neither vμv^{\mu} nor the embedding function of 𝒟1\mathcal{D}_{1} have interesting derivatives. Each metric derivative effectively gives a power of α\alpha. As discussed in section 2.5 around (110), in this setting the fusion metric has only even derivatives on 𝒟1\mathcal{D}_{1}.

Turning now to δΓcusp\delta\Gamma_{\text{cusp}}, it is relatively easy to give a formal expression for it. Equation (B.18a) of Hogervorst:2021spa conveniently encodes the all-order perturbation theory for the ground state energy. Using it, we find

δΓcusp=\displaystyle\delta\Gamma_{\text{cusp}}= 𝒪λ𝒪(α)0|𝒪|0𝒪1𝒪2λ𝒪1(α)λ𝒪2(α)0𝑑τ0|𝒪2(τ)𝒪1(0)|0conn\displaystyle\sum_{{\mathcal{O}}}\lambda_{\mathcal{O}}(\alpha)\langle 0|{\mathcal{O}}|0\rangle-\sum_{{\mathcal{O}}_{1}{\mathcal{O}}_{2}}\lambda_{{\mathcal{O}}_{1}}(\alpha)\lambda_{{\mathcal{O}}_{2}}(\alpha)\int_{0}^{\infty}d\tau\langle 0|{\mathcal{O}}_{2}(\tau){\mathcal{O}}_{1}(0)|0\rangle_{\text{conn}}
+𝒪1𝒪2𝒪3λ𝒪1(α)λ𝒪2(α)λ𝒪3(α)0d2τ0|𝒪3(τ1+τ2)𝒪2(τ1)𝒪1(0)|0conn\displaystyle+\sum_{{\mathcal{O}}_{1}{\mathcal{O}}_{2}{\mathcal{O}}_{3}}\lambda_{{\mathcal{O}}_{1}}(\alpha)\lambda_{{\mathcal{O}}_{2}}(\alpha)\lambda_{{\mathcal{O}}_{3}}(\alpha)\int_{0}^{\infty}d^{2}\tau\langle 0|{\mathcal{O}}_{3}(\tau_{1}+\tau_{2}){\mathcal{O}}_{2}(\tau_{1}){\mathcal{O}}_{1}(0)|0\rangle_{\text{conn}}
,\displaystyle-\cdots, (173)

where τ\tau is the global time on ×Sd1\mathbb{R}\times S^{d-1}. Here, the only dependence on α\alpha is through the couplings λ𝒪(α)\lambda_{\mathcal{O}}(\alpha). Note that the state |0|0\rangle corresponds to an operator on which 𝒟Σ\mathcal{D}_{\Sigma} ends, which is why 0|𝒪|0\langle 0|{\mathcal{O}}|0\rangle is generally non-zero and the two-point functions 0|𝒪2(τ)𝒪1(0)|0\langle 0|{\mathcal{O}}_{2}(\tau){\mathcal{O}}_{1}(0)|0\rangle are not diagonal. If 𝒟Σ\mathcal{D}_{\Sigma} is the trivial defect, the first two orders simplify,

δΓcusp=\displaystyle\delta\Gamma_{\text{cusp}}= 𝒪λ𝒪2(α)0𝑑τ0|𝒪(τ)𝒪(0)|0conn\displaystyle-\sum_{\mathcal{O}}\lambda_{\mathcal{O}}^{2}(\alpha)\int_{0}^{\infty}d\tau\langle 0|{\mathcal{O}}(\tau){\mathcal{O}}(0)|0\rangle_{\text{conn}}
+𝒪1𝒪2𝒪3λ𝒪1(α)λ𝒪2(α)λ𝒪3(α)0d2τ0|𝒪3(τ1+τ2)𝒪2(τ1)𝒪1(0)|0conn\displaystyle+\sum_{{\mathcal{O}}_{1}{\mathcal{O}}_{2}{\mathcal{O}}_{3}}\lambda_{{\mathcal{O}}_{1}}(\alpha)\lambda_{{\mathcal{O}}_{2}}(\alpha)\lambda_{{\mathcal{O}}_{3}}(\alpha)\int_{0}^{\infty}d^{2}\tau\langle 0|{\mathcal{O}}_{3}(\tau_{1}+\tau_{2}){\mathcal{O}}_{2}(\tau_{1}){\mathcal{O}}_{1}(0)|0\rangle_{\text{conn}}
,\displaystyle-\cdots, (174)

where we used that 0|𝒪|0=0\langle 0|{\mathcal{O}}|0\rangle=0 for non-identity operators 𝒪{\mathcal{O}}, and that that the two-point functions are diagonal in this case. Since we are working on the cylinder, the correlation functions in (4.1) and (4.1) are not pure powers and decay exponentially at large times (if they were pure powers, the renormalised integrals would vanish). Finally, note that these expressions contain connected correlators.

One immediate consequence of this is that the small-α\alpha expansion of δΓcusp\delta\Gamma_{\text{cusp}} contains the following powers of α\alpha,

δΓcuspαΔ1++Δnn+2k,\displaystyle\delta\Gamma_{\text{cusp}}\ni\alpha^{\Delta_{1}+\cdots+\Delta_{n}-n+2k}, (175)

where n1n\geq 1 and k0k\geq 0 are integers, while Δi\Delta_{i} run over the scaling dimension of the primary operators in (168). In particular, δΓcusp=O(αΔmin1)\delta\Gamma_{\text{cusp}}=O(\alpha^{\Delta_{\text{min}}-1}), where Δmin\Delta_{\text{min}} is the smallest scaling dimension of an operator appearing in (168).38{}^{\ref{footnote:marginal}}. When 𝒟Σ\mathcal{D}_{\Sigma} is trivial, this is modified in the obvious way and we find that δΓcusp=O(α2Δmin2)\delta\Gamma_{\text{cusp}}=O(\alpha^{2\Delta_{\text{min}}-2}).

The above discussion can be straightforwardly generalised to excited states, although (4.1) needs to be modified in that case Hogervorst:2021spa . One important difference is that one-point functions generically do not vanish in excited states even for trivial 𝒟Σ\mathcal{D}_{\Sigma}.

4.2 Supersymmetric Wilson lines in 𝒩=4\mathcal{N}=4 SYM

In this section we briefly examine the fusion of fundamental and anti-fundamental supersymmetric Wilson lines Maldacena:1998im in 𝒩=4\mathcal{N}=4 SYM, focusing on the planar theory.

A (locally) supersymmetric Wilson line along a curve 𝒟\mathcal{D} is defined by

Pei𝒟A+𝒟𝑑s(nϕ),\displaystyle Pe^{i\int_{\mathcal{D}}A+\int_{\mathcal{D}}ds(n\cdot\phi)}, (176)

where ϕ\phi is the fundamental scalar and n6n\in\mathbb{R}^{6} is a unit vector in the vector representation of SO(6)\mathrm{SO}(6) R-symmetry (which is therefore explicitly broken to SO(5)\mathrm{SO}(5) by the Wilson line). The path-ordered exponential is computed in the fundamental representation of the SU(N)\mathrm{SU}(N) gauge group. We will denote the resulting defect by 𝒟n\mathcal{D}_{n}.

We consider the fusion of 𝒟1=𝒟n1\mathcal{D}_{1}=\mathcal{D}_{n_{1}} with 𝒟2=𝒟¯n2\mathcal{D}_{2}=\overline{\mathcal{D}}_{n_{2}} and we use θ\theta to denote the angle between n1n_{1} and n2n_{2}. We expect that the resulting IR defect 𝒟Σ\mathcal{D}_{\Sigma} is the trivial defect. Informally, the pair of quarks represented by the Wilson lines can form a singlet or an adjoint state, and the former is more energetically favorable.393939The potential in the adjoint representation vanishes in the planar theory Pineda:2007kz , while the singlet potential is non-zero and attractive. More formally, the line 𝒟Σ\mathcal{D}_{\Sigma} can end and we will show that the spectrum of local operators it can end on coincides with the bulk spectrum. In other words, we will describe the spectrum of the Hamiltonian HΣH_{\Sigma} on ×S3\mathbb{R}\times S^{3} when there is a single insertion of 𝒟Σ\mathcal{D}_{\Sigma} along the time direction \mathbb{R}, and we shall see that it agrees with the spectrum of the bulk dilatation operator.

To this end, we consider the configuration discussed in section 4.1 where the straight lines 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} meet at a cusp with an opening angle α=πϕ\alpha=\pi-\phi. We also retain the notation from section 4.1. Due to (168), (170) and (172), the spectrum of HΣH_{\Sigma} is obtained from the spectrum of HH in the limit α0\alpha\to 0 after subtracting the singular part of ground state energy E0=ΓcuspE_{0}=\Gamma_{\text{cusp}}. We will see in various limits below that Γcusp=a0α+O(α)\Gamma_{\text{cusp}}=-\frac{a_{0}}{\alpha}+O(\alpha), and we expect this to hold non-perturbatively.404040Nikolay Gromov has informed us that unpublished numerical data for Γcusp(α)\Gamma_{\text{cusp}}(\alpha) at g=λ/4π=3/8g=\sqrt{\lambda}/4\pi=3/8 and θ=0\theta=0 obtained from QSC in Gromov:2015dfa shows that the coefficient of α0\alpha^{0} in Γcusp\Gamma_{\text{cusp}} is 0±1040\pm 10^{-4}. By comparison, a00.80a_{0}\approx 0.80 and the coefficient of α\alpha is 0.23\approx 0.23. Therefore, we can simply subtract the full Γcusp\Gamma_{\text{cusp}}.

We can classify the cusp operators into single-trace and multi-trace. The single-trace operators are obtained from insertions of fundamental fields in the Wilson line trace, while multi-trace operators may have additional traces. An example of a single-trace operator is

Tr{Pei𝒟1A+𝒟1𝑑s(n1ϕ)ϕiPei𝒟2A+𝒟2𝑑s(n2ϕ)},\displaystyle\mathrm{Tr}\{\cdots Pe^{i\int_{\mathcal{D}_{1}}A+\int_{\mathcal{D}_{1}}ds(n_{1}\cdot\phi)}\phi_{i}Pe^{i\int_{\mathcal{D}_{2}}A+\int_{\mathcal{D}_{2}}ds(n_{2}\cdot\phi)}\cdots\}, (177)

while an example of a multi-trace operator is

Tr(ϕiϕi)Tr{Pei𝒟1A+𝒟1𝑑s(n1ϕ)Pei𝒟2A+𝒟2𝑑s(n2ϕ)},\displaystyle\mathrm{Tr}(\phi_{i}\phi_{i})\mathrm{Tr}\{\cdots Pe^{i\int_{\mathcal{D}_{1}}A+\int_{\mathcal{D}_{1}}ds(n_{1}\cdot\phi)}Pe^{i\int_{\mathcal{D}_{2}}A+\int_{\mathcal{D}_{2}}ds(n_{2}\cdot\phi)}\cdots\}, (178)

where Tr(ϕiϕi)\mathrm{Tr}(\phi_{i}\phi_{i}) is inserted at the cusp. Note that in the planar theory the scaling dimension of (178) is just Γcusp+ΔK=E0+ΔK\Gamma_{\text{cusp}}+\Delta_{K}=E_{0}+\Delta_{K} where ΔK\Delta_{K} is the scaling dimension of the bulk Konishi operator Tr(ϕiϕi)\mathrm{Tr}(\phi_{i}\phi_{i}). Therefore, the spectrum of HΣH_{\Sigma} contains ΔK\Delta_{K}. In a similar way, it contains the scaling dimension of any bulk single- or multi-trace operator. On the other hand, we expect that the excited single-trace states such as (177) have energies which behave as Ea0αE\sim-\frac{a^{\prime}_{0}}{\alpha} with a0<a0a^{\prime}_{0}<a_{0}.414141Note that this is a non-degeneracy assumption on a0a_{0}. If there were an exact degeneracy for generic couplings, it would also be present in the ladder limit. We will see in section 4.3 that in the ladder limit a0a_{0} is non-degenerate. Therefore, EE0+E-E_{0}\to+\infty as α0\alpha\to 0 for states which involve such single-trace factors, and only the multi-trace operators similar to (178) survive in the spectrum of HΣH_{\Sigma}. This shows that the spectrum of HΣH_{\Sigma} agrees with the spectrum of the bulk dilatation operator and supports our claim that 𝒟Σ\mathcal{D}_{\Sigma} is trivial.

If we focus on a finite set of lowest-energy states of HH, then the above argument shows that for sufficiently small α\alpha their energies are given exactly by

H=low-E spectrumΓcusp+HΣ.\displaystyle H\overset{\text{low-E spectrum}}{=}\Gamma_{\text{cusp}}+H_{\Sigma}. (179)

In particular, the α\alpha-dependence of HH energy levels is state-independent. This statement should have non-trivial implications for the couplings λ𝒪(α)\lambda_{\mathcal{O}}(\alpha) entering in (168). We leave this question for the future. Here we will focus on the contribution λ1(α)\lambda_{\textbf{1}}(\alpha) of the identity operator. Since all non-trivial bulk local operators in 𝒩=4\mathcal{N}=4 SYM have dimension at least 2, it follows that, at least, Γcusp=λ1(α)+O(α2)\Gamma_{\text{cusp}}=\lambda_{\textbf{1}}(\alpha)+O(\alpha^{2}).

We therefore find for the cusp anomalous dimension, taking into account (172),

Γcusp(θ,ϕ=πα)=a0(θ)α3a2,2(θ)α+O(α2),\displaystyle\Gamma_{\text{cusp}}(\theta,\phi=\pi-\alpha)=-\frac{a_{0}(\theta)}{\alpha}-3a_{2,2}(\theta)\alpha+O(\alpha^{2}), (180)

where a2,2a_{2,2} is the Wilson coefficient appearing in (171). The function Γcusp(θ,ϕ)\Gamma_{\text{cusp}}(\theta,\phi) can in principle be computed using integrability Gromov:2015dfa , and therefore the Wilson coefficients a0a_{0}, a2,2a_{2,2}, as well as other coefficients determining subleading terms in Γcusp\Gamma_{\text{cusp}} are accessible through integrability techniques. The coefficient a0a_{0} has been studied using Quantum Spectral Curve (QSC) in Gromov:2016rrp . It would be interesting to find a QSC description for the subleading coefficients such as a2,2a_{2,2}. Numerical data from QSC supports the absence of α0\alpha^{0} term, see footnote 40. In the following two subsections we address a simpler problem and compute a2,2a_{2,2} in the ladder limit and in the strong coupling limit.

4.3 Ladder limit

The first limit we consider is the limit eiθe^{i\theta}\to\infty and ’t Hooft coupling λ0\lambda\to 0 with

λ^=λeiθ4\displaystyle\widehat{\lambda}=\frac{\lambda e^{i\theta}}{4} (181)

held fixed. As discussed in Correa:2012nk , in this limit only the ladder diagrams contribute to Γcusp\Gamma_{\text{cusp}}, which can then be determined through the ground state energy of the Schrödinger equation Correa:2012nk

ψ′′(x)λ^8π21coshxcosαψ(x)=Γcusp24ψ(x),\displaystyle-\psi^{\prime\prime}(x)-\frac{\widehat{\lambda}}{8\pi^{2}}\frac{1}{\cosh x-\cos\alpha}\psi(x)=-\frac{\Gamma_{\text{cusp}}^{2}}{4}\psi(x), (182)

where xx\in\mathbb{R}.

We are interested in the limit α0\alpha\to 0. In this limit, the potential becomes singular at x=0x=0 and Γcusp\Gamma_{\text{cusp}} behaves according to (180). To study this limit, we write x=αyx=\alpha y and

Γcusp=Ωα.\displaystyle\Gamma_{\text{cusp}}=-\frac{\Omega}{\alpha}. (183)

Equation (182) becomes

ψ′′(y)λ^8π2α2coshαycosαψ(y)=Ω24ψ(y).\displaystyle-\psi^{\prime\prime}(y)-\frac{\widehat{\lambda}}{8\pi^{2}}\frac{\alpha^{2}}{\cosh\alpha y-\cos\alpha}\psi(y)=-\frac{\Omega^{2}}{4}\psi(y). (184)

The potential now has a regular expansion at small α\alpha,

λ^8π2α2coshαycosα=U0(y)+α2U2(y)+,\displaystyle-\frac{\widehat{\lambda}}{8\pi^{2}}\frac{\alpha^{2}}{\cosh\alpha y-\cos\alpha}=U_{0}(y)+\alpha^{2}U_{2}(y)+\cdots, (185)

where

U0(y)=λ^4π211+y2,U2(y)=16U0(y)+λ^48π2.\displaystyle U_{0}(y)=-\frac{\widehat{\lambda}}{4\pi^{2}}\frac{1}{1+y^{2}},\quad U_{2}(y)=\tfrac{1}{6}U_{0}(y)+\frac{\widehat{\lambda}}{48\pi^{2}}. (186)

Let Ω02(λ^)/4-\Omega_{0}^{2}(\widehat{\lambda})/4 be the ground state energy of the leading-order Schrödinger equation

ψ′′(y)λ^4π211+y2ψ(y)=Ω02(λ^)4ψ(y).\displaystyle-\psi^{\prime\prime}(y)-\frac{\widehat{\lambda}}{4\pi^{2}}\frac{1}{1+y^{2}}\psi(y)=-\frac{\Omega_{0}^{2}(\widehat{\lambda})}{4}\psi(y). (187)

Equation (185) then implies that

Ω2/4=Ω02(λ^(1+α26))/4+λ^α248π2+O(α4).\displaystyle-\Omega^{2}/4=-\Omega_{0}^{2}\left(\widehat{\lambda}(1+\tfrac{\alpha^{2}}{6})\right)/4+\frac{\widehat{\lambda}\alpha^{2}}{48\pi^{2}}+O(\alpha^{4}). (188)

In other words,

Ω=Ω0(λ^)14π2Ω0(λ^)λ^Ω0(λ^)6Ω0(λ^)/λ^α2+O(α4)\displaystyle\Omega=\Omega_{0}(\widehat{\lambda})-\frac{\tfrac{1}{4\pi^{2}}-\Omega_{0}(\widehat{\lambda})\partial_{\widehat{\lambda}}\Omega_{0}(\widehat{\lambda})}{6\Omega_{0}(\widehat{\lambda})/\widehat{\lambda}}\alpha^{2}+O(\alpha^{4}) (189)

Using (183) and (180) this gives for the Wilson coefficients

a0=Ω0(λ^),a2,2=Ω0(λ^)λ^Ω0(λ^)14π218Ω0(λ^)/λ^.\displaystyle a_{0}=\Omega_{0}(\widehat{\lambda}),\quad a_{2,2}=\frac{\Omega_{0}(\widehat{\lambda})\partial_{\widehat{\lambda}}\Omega_{0}(\widehat{\lambda})-\tfrac{1}{4\pi^{2}}}{18\Omega_{0}(\widehat{\lambda})/\widehat{\lambda}}. (190)

The weak-coupling expansion of Ω0(λ^)\Omega_{0}(\widehat{\lambda}) was computed in Correa:2012nk to O(λ^3)O(\widehat{\lambda}^{3}). For simplicity, we reproduce the first two orders,

a0=Ω0(λ^)=λ^4π+λ^28π3(logλ^2π+γE1)+O(λ^3).\displaystyle a_{0}=\Omega_{0}(\widehat{\lambda})=\frac{\widehat{\lambda}}{4\pi}+\frac{\widehat{\lambda}^{2}}{8\pi^{3}}\left(\log\tfrac{\widehat{\lambda}}{2\pi}+\gamma_{E}-1\right)+O(\widehat{\lambda}^{3}). (191)

This yields

a2,2=118π+λ^36π3(logλ^2π+γE1+π22)+O(λ^2).\displaystyle a_{2,2}=-\frac{1}{18\pi}+\frac{\widehat{\lambda}}{36\pi^{3}}\left(\log\tfrac{\widehat{\lambda}}{2\pi}+\gamma_{E}-1+\tfrac{\pi^{2}}{2}\right)+O(\widehat{\lambda}^{2}). (192)

It may seem surprising that a2,2a_{2,2} starts at O(1)O(1) while at fixed α\alpha the expansion of Γcusp\Gamma_{\text{cusp}} starts at O(λ^)O(\widehat{\lambda}). This, as well as the presence of the logarithms of the coupling, is due to the fact that the small-α\alpha expansion of Γcusp\Gamma_{\text{cusp}} does not commute with the small-λ^\widehat{\lambda} (more generally, small λ\lambda) expansion Correa:2012nk ; Pineda:2007kz . In particular, the expansion (180) is only valid for angles αλ^\alpha\ll\widehat{\lambda}. We briefly discuss the origin of this in section 3.5 from the fusion point of view.

Finally, we note that the ground state of (187) is non-degenerate and therefore the first excited energy of (182) scales as a0/α-a_{0}^{\prime}/\alpha with a0<a0a_{0}^{\prime}<a_{0}. These energies correspond to other single-trace insertions at the cusp.424242Equation (182) describes a subset of single-trace insertions at the cusp (as bound states or resonances, see Cavaglia:2018lxi ). The remaining single-trace states have protected scaling dimensions in the ladder limit, i.e. for them a0=0a^{\prime}_{0}=0. Still, a0<a0a^{\prime}_{0}<a_{0} is satisfied. We thank Nikolay Gromov for discussions on this point. This supports the claim that we made earlier in this section.

4.4 Strong coupling

It is also possible to compute a0a_{0} and a2,2a_{2,2} at strong coupling using the results of Drukker:2011za . In particular, Drukker:2011za gives an explicit expression for the leading O(λ)O(\sqrt{\lambda}) term in Γcusp\Gamma_{\text{cusp}} for generic α\alpha and ϕ\phi in their appendix B, and they also extract the leading coefficient a0a_{0}. By a straightforward extension of their analysis we find

a0(θ)\displaystyle a_{0}(\theta) =2λk1k2π(E(1k2)K)2+O(λ0),\displaystyle=\frac{2\sqrt{\lambda}}{k\sqrt{1-k^{2}}\pi}\left(E-(1-k^{2})K\right)^{2}+O(\lambda^{0}),
a2,2(θ)\displaystyle a_{2,2}(\theta) =λ36k1k2π(12k2)2E2(1k2)(3k44k2+2)EK+(1k2)2K2(E(1k2)K)((12k2)E(1k2)K)+O(λ0),\displaystyle=-\frac{\sqrt{\lambda}}{36k\sqrt{1-k^{2}}\pi}\frac{(1-2k^{2})^{2}E^{2}-(1-k^{2})(3k^{4}-4k^{2}+2)EK+(1-k^{2})^{2}K^{2}}{\left(E-(1-k^{2})K\right)\left((1-2k^{2})E-(1-k^{2})K\right)}+O(\lambda^{0}), (193)

where K=K(k2)K=K(k^{2}) and E=E(k2)E=E(k^{2}) are complete elliptic integrals of the first and second kind, respectively. The elliptic modulus kk is related to θ\theta via

θ=212k2K(k2).\displaystyle\theta=2\sqrt{1-2k^{2}}K(k^{2}). (194)

In the special case of θ=0\theta=0 we obtain

a0(0)=4π2λΓ(14)4+O(λ0),a2,2(0)=λ(24π2Γ(14)4)576π3+O(λ0).\displaystyle a_{0}(0)=\frac{4\pi^{2}\sqrt{\lambda}}{\Gamma(\tfrac{1}{4})^{4}}+O(\lambda^{0}),\quad a_{2,2}(0)=-\frac{\sqrt{\lambda}\left(24\pi^{2}-\Gamma(\tfrac{1}{4})^{4}\right)}{576\pi^{3}}+O(\lambda^{0}). (195)

We note in passing that a0(0)>0a_{0}(0)>0 and is the global maximum of a0(θ)a_{0}(\theta) as required by theorems 3.1 and 3.2.

5 Local operator one-point function asymptotics

In this section we study the implications of conformal defect fusion for the OPE of the defect two-point function 𝒟1𝒟2\langle\mathcal{D}_{1}\mathcal{D}_{2}\rangle. Recall that each of the two defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} can be expanded in a basis of local operators Gadde:2016fbj , which yields a convergent expansion of the schematic form

𝒟1𝒟2=𝒪C𝒟1𝒪C𝒟2𝒪G𝒪.\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\rangle=\sum_{\mathcal{O}}C_{\mathcal{D}_{1}{\mathcal{O}}}C_{\mathcal{D}_{2}{\mathcal{O}}^{\dagger}}G_{{\mathcal{O}}}. (196)

Here, the sum is over the primary operators 𝒪{\mathcal{O}}, C𝒟i𝒪C_{\mathcal{D}_{i}{\mathcal{O}}} is the one-point function of 𝒪{\mathcal{O}} in the presence of 𝒟i\mathcal{D}_{i}, and G𝒪G_{{\mathcal{O}}} is a conformal block which depends only on the cross-ratios describing the configuration of 𝒟i\mathcal{D}_{i} and on the quantum numbers of 𝒪{\mathcal{O}}. Compatibility of this OPE with the fusion (2) leads to relations between the Wilson coefficients in the effective action and the one-point functions C𝒟1𝒪C_{\mathcal{D}_{1}{\mathcal{O}}} of primary operators. This is the relation that we will explore.

We will focus on the case 𝒟1𝒟¯2\mathcal{D}_{1}\simeq\overline{\mathcal{D}}_{2} in which case C𝒟1𝒪C𝒟2𝒪=|C𝒟1𝒪|20C_{\mathcal{D}_{1}{\mathcal{O}}}C_{\mathcal{D}_{2}{\mathcal{O}}^{\dagger}}=|C_{\mathcal{D}_{1}{\mathcal{O}}}|^{2}\geq 0. Furthermore, we will simplify the problem by studying a simpler expansion in which the descendants and the primaries are treated independently.

5.1 Defect two-point function from OPE expansion

We consider the generic configuration of two dimension-pp conformal defects that we previously studied as an example in section 2.1. Specifically, we take 𝒟1\mathcal{D}_{1} to be a pp-dimensional spherical defect of radius 11 centred at 0 and lying in the subspace spanned by the coordinate directions 1,,p+11,\cdots,p+1. The position of the defect 𝒟2\mathcal{D}_{2} is obtained by the action of

elogr𝐝iθi𝐦i,p+i+1\displaystyle e^{\log r\mathbf{d}-\sum_{i}\theta_{i}\mathbf{m}_{i,p+i+1}} (197)

where the sum is over i=1,,m=min{p+1,dp1}i=1,\cdots,m=\min\{p+1,d-p-1\}. The total number of parameters r,θ1,,θmr,\theta_{1},\cdots,\theta_{m} is m+1=min{d+2q,q}m+1=\min\{d+2-q,q\}. This is the same as the number of cross-ratios for a pair of pp-dimensional defects Gadde:2016fbj , and thus the configurations of the above form cover at least an open neighbourhood of the fusion limit r=1,θi=0r=1,\theta_{i}=0 in the cross-ratio space.

When 𝒟2\mathcal{D}_{2} is conjugate to 𝒟1\mathcal{D}_{1}, we can write the two-point correlation function as

𝒟1𝒟2=𝒟¯1|elogrDiθiMi,p+i+1|𝒟¯1=n𝒟¯1|nn|elogrDiθiMi,p+i+1|𝒟¯1,\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\rangle=\langle\overline{\mathcal{D}}_{1}|e^{\log rD-\sum_{i}\theta_{i}M_{i,p+i+1}}|\overline{\mathcal{D}}_{1}\rangle=\sum_{n}\langle\overline{\mathcal{D}}_{1}|n\rangle\langle n|e^{\log rD-\sum_{i}\theta_{i}M_{i,p+i+1}}|\overline{\mathcal{D}}_{1}\rangle, (198)

where all states are in the Hilbert space of the unit sphere, and the sum over nn is a sum over an orthonormal basis of states. Since DD and Mi,p+i+1M_{i,p+i+1} with i=1,,mi=1,\cdots,m are mutually commuting, we can assume that they are diagonalised in our basis,

D|n\displaystyle D|n\rangle =Δn|n,\displaystyle=\Delta_{n}|n\rangle, (199)
Mi,p+i+1|n\displaystyle M_{i,p+i+1}|n\rangle =iJi,n|n.\displaystyle=iJ_{i,n}|n\rangle. (200)

The expression for the two-point function then becomes

𝒟1𝒟2=n|𝒟¯1|n|2rΔneiiJi,nθi.\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\rangle=\sum_{n}|\langle\overline{\mathcal{D}}_{1}|n\rangle|^{2}r^{\Delta_{n}}e^{-i\sum_{i}J_{i,n}\theta_{i}}. (201)

Notice that 𝒟¯1|n\langle\overline{\mathcal{D}}_{1}|n\rangle is equal to the one-point function of the local operator corresponding to the state |n|n\rangle in the presence of the defect 𝒟1\mathcal{D}_{1}.

We can rewrite the above equation equivalently as

𝒟1𝒟2=𝑑Δ𝑑J1𝑑Jmρ(Δ,J1,,Jm)rΔeiiJiθi,\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\rangle=\int d\Delta dJ_{1}\cdots dJ_{m}\rho(\Delta,J_{1},\cdots,J_{m})r^{\Delta}e^{-i\sum_{i}J_{i}\theta_{i}}, (202)

where

ρ(Δ,J1,,Jm)=n|𝒟¯1|n|2δ(ΔΔn)δ(J1J1,n)δ(JmJm,n)\displaystyle\rho(\Delta,J_{1},\cdots,J_{m})=\sum_{n}|\langle\overline{\mathcal{D}}_{1}|n\rangle|^{2}\delta(\Delta-\Delta_{n})\delta(J_{1}-J_{1,n})\cdots\delta(J_{m}-J_{m,n}) (203)

is a non-negative density of one-point functions.

The density ρ(Δ,J1,,Jm)\rho(\Delta,J_{1},\cdots,J_{m}) receives contributions from both primaries and descendants. It is possible to recover the density of one-point functions of primary operators from ρ\rho by analyzing the series expansion of the conformal block G𝒪G_{\mathcal{O}}. In this context, it is important to note that only the primary operators in spin representations described by at most mm-row Young diagrams can have non-zero one-point functions with a pp-dimensional defect Lauria:2018klo . Therefore, ρ(Δ,J1,,Jm)\rho(\Delta,J_{1},\cdots,J_{m}) has the correct number of variables to discern all non-zero one-point functions of primaries.

We will focus on the density ρ(Δ,J1,,Jm)\rho(\Delta,J_{1},\cdots,J_{m}). We will see that, within the precision that we work at, the density of primary one-point functions has the same form.

5.2 Defect two-point function from fusion

The limit r1r\to 1 and θi0\theta_{i}\to 0 corresponds to the fusion of the defects 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}, allowing us to obtain an approximation to the two-point function 𝒟1𝒟2\langle\mathcal{D}_{1}\mathcal{D}_{2}\rangle using (2). Specifically, we have the following equality in the L0L\to 0 limit

𝒟1𝒟2𝒟Σ[eSeff]\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\rangle\sim\langle\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}}]\rangle (204)

with r=1+O(L),θi=O(L)r=1+O(L),\theta_{i}=O(L). Note that the right-hand side should be understood as an asymptotic series in the obvious way. In particular, this does not give an exact prediction for the two-point function at any finite value of LL.

Let us split Seff=Seff𝟏+SeffrestS_{\text{eff}}=S^{\mathbf{1}}_{\text{eff}}+S^{\text{rest}}_{\text{eff}}, where Seff𝟏S^{\mathbf{1}}_{\text{eff}} contains the contribution of the identity operator, and SeffrestS^{\text{rest}}_{\text{eff}} contains the contributions from all the other operators, which are all irrelevant.434343See section 3 for a more precise discussion, in particular relating to marginal operators. We then have

𝒟Σ[eSeff]=eSeff𝟏𝒟Σ[eSeffrest]=eSeff𝟏(𝒟Σ+o(1)).\displaystyle\langle\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}}]\rangle=e^{-S^{\mathbf{1}}_{\text{eff}}}\langle\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}^{\text{rest}}}]\rangle=e^{-S^{\mathbf{1}}_{\text{eff}}}\left(\langle\mathcal{D}_{\Sigma}\rangle+o(1)\right). (205)

Here the o(1)o(1) terms go to 0 as L0L\to 0 and are given by conformal perturbation theory in terms of integrated correlation functions of irrelevant operators on 𝒟Σ\mathcal{D}_{\Sigma}. For example, the leading contribution is

12dpz1γ^dpz2γ^λ𝒪^(z1)λ𝒪^(z2)𝒟Σ[𝒪^(z1)𝒪^(z2)]=O(L2Δ𝒪2p),\displaystyle\frac{1}{2}\int d^{p}z_{1}\sqrt{\widehat{\gamma}}d^{p}z_{2}\sqrt{\widehat{\gamma}}\lambda_{\widehat{{\mathcal{O}}}}(z_{1})\lambda_{\widehat{{\mathcal{O}}}}(z_{2})\langle\mathcal{D}_{\Sigma}[\widehat{{\mathcal{O}}}(z_{1})\widehat{{\mathcal{O}}}(z_{2})]\rangle=O(L^{2\Delta_{\mathcal{O}}-2p}), (206)

where 𝒪{\mathcal{O}} is the leading irrelevant operator on 𝒟Σ\mathcal{D}_{\Sigma} (assuming it is Hermitian and unique).

We will only study the implications of the leading cosmological constant term (129) in Seff𝟏S_{\text{eff}}^{\mathbf{1}}. That is, we will work with the leading approximation

𝒟1𝒟2\displaystyle\langle\mathcal{D}_{1}\mathcal{D}_{2}\rangle =exp{Seff(0)+O(Lp+1)}(𝒟Σ+O(L2Δ𝒪2p))\displaystyle=\exp\left\{-S^{(0)}_{\text{eff}}+O(L^{-p+1})\right\}\left(\langle\mathcal{D}_{\Sigma}\rangle+O(L^{2\Delta_{\mathcal{O}}-2p})\right)
=exp{a0dpzγp(z)+O(Lp+1)}(𝒟Σ+O(L2Δ𝒪2p)).\displaystyle=\exp\left\{a_{0}\int d^{p}z\sqrt{\gamma}\ell^{-p}(z)+O(L^{-p+1})\right\}\left(\langle\mathcal{D}_{\Sigma}\rangle+O(L^{2\Delta_{\mathcal{O}}-2p})\right). (207)

It remains to evaluate the leading term (129) in the fusion limit in our kinematics. The relevant scale function (z)\ell(z) was computed in section 2.2, see (50). Let us parameterise the fusion limit as

r=eL,θi=LΩi,\displaystyle r=e^{-L},\hskip 28.45274pt\theta_{i}=L\Omega_{i}, (208)

where L0L\to 0 is now a concrete parameter defined by the above equations. We obtain

Seff(0)\displaystyle S_{\text{eff}}^{(0)} =a0dpzγp(z)=a0Lpdpzγ1(1+i=1mΩi2xi2)p/2,\displaystyle=-a_{0}\int d^{p}z\sqrt{\gamma}~{}\ell^{-p}(z)=-\frac{a_{0}}{L^{p}}\int d^{p}z\sqrt{\gamma}\frac{1}{\left(1+\sum_{i=1}^{m}\Omega_{i}^{2}x_{i}^{2}\right)^{p/2}}, (209)

where xix_{i} are the bulk coordinates of the point zz on the defect 𝒟1\mathcal{D}_{1}. In the simplest case, Ωi=0\Omega_{i}=0 and we find

Seff(0)\displaystyle S_{\text{eff}}^{(0)} =a0dpzγp(z)=a0volSpLp,\displaystyle=-a_{0}\int d^{p}z\sqrt{\gamma}~{}\ell^{-p}(z)=-\frac{a_{0}\mathop{\mathrm{vol}}S^{p}}{L^{p}}, (210)

where volSp\mathop{\mathrm{vol}}S^{p} is the volume of unit pp-sphere, i.e. volSp=2π(p+1)/2Γ(p+12)\mathop{\mathrm{vol}}S^{p}=\frac{2\pi^{(p+1)/2}}{\Gamma\left(\frac{p+1}{2}\right)}.

For non-zero Ωi\Omega_{i} the integration over dpzd^{p}z seems to be hard to carry out in full generality, so we focus on the case of p=1p=1. In this case m=min{p+1,dp1}=2m=\min\{p+1,d-p-1\}=2 for d4d\geq 4 and m=1m=1 in d=3d=3. In general, the integral to perform is

Seff(0)\displaystyle S_{\text{eff}}^{(0)} =a0L02π𝑑ϕ11+Ω12cos2ϕ+Ω22sin2ϕ=4a0L1+Ω22K(Ω12+Ω221+Ω22),\displaystyle=-\frac{a_{0}}{L}\int_{0}^{2\pi}d\phi\frac{1}{\sqrt{1+\Omega_{1}^{2}\cos^{2}\phi+\Omega_{2}^{2}\sin^{2}\phi}}=-\frac{4a_{0}}{L\sqrt{1+\Omega_{2}^{2}}}K\left(\frac{-\Omega_{1}^{2}+\Omega_{2}^{2}}{1+\Omega_{2}^{2}}\right), (211)

where K(z)K(z) is the complete elliptic integral of the first kind. Note that the right-hand side is symmetric in Ω1,Ω2\Omega_{1},\Omega_{2} despite the appearance.444444This follows from the identity K(z)=11zK(zz1)K(z)=\tfrac{1}{\sqrt{1-z}}K(\tfrac{z}{z-1}). In d=3d=3 we have to set Ω2=0\Omega_{2}=0 in the above equation.

5.3 Spinless one-point function density for general defects

The consistency of the OPE (202) and the fusion (5.2) results for the defect two-point function leads to the equation

𝑑Δ𝑑J1𝑑Jmρ(Δ,J1,,Jm)rΔeiiJiθiexp{Seff(0)+O(Lp+1)}.\displaystyle\int d\Delta dJ_{1}\cdots dJ_{m}\rho(\Delta,J_{1},\cdots,J_{m})r^{\Delta}e^{-i\sum_{i}J_{i}\theta_{i}}\sim\exp\left\{-S^{(0)}_{\text{eff}}+O(L^{-p+1})\right\}. (212)

We will now study the implications of this equation, starting with the case Ωi=0\Omega_{i}=0 in (208).

In this case θi=0\theta_{i}=0 and the left-hand side of (212) can be rewritten as

𝑑Δρ(Δ)rΔ\displaystyle\int d\Delta\rho(\Delta)r^{\Delta} (213)

where

ρ(Δ)=𝑑J1𝑑Jmρ(Δ,J1,,Jm)\displaystyle\rho(\Delta)=\int dJ_{1}\cdots dJ_{m}\rho(\Delta,J_{1},\cdots,J_{m}) (214)

is the total density of defect one-point functions at scaling dimension Δ\Delta. The right-hand side of (212) can be evaluated using (210), giving

𝑑Δρ(Δ)rΔ=exp{a0volSp(1r)p+O((1r)p+1)}.\displaystyle\int d\Delta\rho(\Delta)r^{\Delta}=\exp\left\{\frac{a_{0}\mathop{\mathrm{vol}}S^{p}}{(1-r)^{p}}+O((1-r)^{-p+1})\right\}. (215)

It is well-known Cardy:1986ie ; Pappadopulo:2012jk that such relations constrain the large-Δ\Delta behaviour of ρ(Δ)\rho(\Delta). The simple way to extract the density ρ(Δ)\rho(\Delta) from the above expression is using an inverse Laplace transform,

ρ(Δ)=12πiy0iy0+i𝑑yexp{a0volSp(1ey)p+O(yp+1)}eΔy,\displaystyle\rho(\Delta)=\frac{1}{2\pi i}\int_{y_{0}-i\infty}^{y_{0}+i\infty}\,dy\exp\left\{\frac{a_{0}\mathop{\mathrm{vol}}S^{p}}{(1-e^{-y})^{p}}+O(y^{-p+1})\right\}~{}e^{\Delta y}, (216)

where y0y_{0} has to be chosen sufficiently large real part so that all the singularities of the integrand are to the left of the integration contour. An approximation to ρ(Δ)\rho(\Delta) can be obtained under the assumption that for Δ1\Delta\gg 1 the integral is dominated by a saddle point at small yy where we have an approximation for the integrand.

Working under this assumption, we find the leading-order saddle-point equation at large Δ\Delta (recall that a00a_{0}\geq 0)

ya0volSpyp+Δ=0y(pa0volSpΔ)1p+1,\displaystyle\frac{\partial}{\partial y}\frac{a_{0}\mathop{\mathrm{vol}}S^{p}}{y^{p}}+\Delta=0\hskip 8.5359pt\implies\hskip 8.5359pty_{*}\approx\left(\frac{pa_{0}\mathop{\mathrm{vol}}S^{p}}{\Delta}\right)^{\frac{1}{p+1}}, (217)

which yields for the leading one-point function density

ρ(Δ)exp[(1+p)(a0volSp)1p+1(Δp)pp+1].\displaystyle\rho(\Delta)\sim\exp\left[(1+p)\left(a_{0}\mathop{\mathrm{vol}}S^{p}\right)^{\frac{1}{p+1}}\left(\frac{\Delta}{p}\right)^{\frac{p}{p+1}}\right]. (218)

This can be compared with the asymptotic density of states in a dd-dimensional CFT,

ρstates(Δ)exp[hΔd1d]\displaystyle\rho_{\text{states}}(\Delta)\sim\exp\left[h\Delta^{\frac{d-1}{d}}\right] (219)

where h>0h>0 can be expressed in terms of free energy density Benjamin:2023qsc . For p<d1p<d-1 we have Δpp+1Δd1d\Delta^{\frac{p}{p+1}}\ll\Delta^{\frac{d-1}{d}}, and thus on average the one-point functions of local operators in the presence of codimension-qq defect with q>1q>1 have to be very small.

Equation (218) gives the one-point function density for all operators including descendants. The density of primaries has the same leading term (218) since we expect the series coefficients in the rr-expansion of the conformal blocks G𝒪G_{\mathcal{O}} appearing in (196) to grow only polynomially with degree.454545Note however that we did not prove this in full generality.

The result (218) was derived under some assumptions on the behaviour of the inverse Laplace transform. This approach, although not fully justified, can be taken further to compute subleading corrections to (218) by a more careful evaluation of the saddle point integral and taking into account subleading terms in the fusion effective action. Alternatively, results of this kind can be obtained rigorously using various Tauberian theorems Pappadopulo:2012jk ; Qiao:2017xif ; Mukhametzhanov:2018zja ; Mukhametzhanov:2019pzy , although the subleading terms are hard to control and depend on the coarse-graining prescription Mukhametzhanov:2019pzy .464646Coarse-graining is necessary to interpret (218) since the left-hand side is a sum of delta-functions.

Ignoring these subtleties, let us examine the first subleading term in (218). It can come from the saddle-point expansion of the integral or from the subleading terms in the identity coupling in SeffS_{\text{eff}}. The latter start at two-derivative order (the one-derivative term (147) vanishes in these kinematics). Therefore, they are O(1)O(1) or smaller for p2p\leq 2. On the other hand, the saddle-point integral will in general produce logΔ\log\Delta terms in the exponent. We thus find the leading corrections for p=1,2p=1,2,

ρp=1(Δ)\displaystyle\rho_{p=1}(\Delta) Δ3/4exp[22πa0Δ1/2],\displaystyle\sim\Delta^{-3/4}\exp\left[2\sqrt{2\pi a_{0}}\Delta^{1/2}\right],
ρp=2(Δ)\displaystyle\rho_{p=2}(\Delta) Δ2/3exp[3(πa0)1/3Δ2/3].\displaystyle\sim\Delta^{-2/3}\exp\left[3\left(\pi a_{0}\right)^{1/3}\Delta^{2/3}\right]. (220)

Further subleading corrections can be straightforwardly incorporated. For p>2p>2 the leading corrections come from the 2-derivative terms in the identity coupling.

5.4 Spin-dependent one-point function density for line defects

We now consider the more general situation when Ωi0\Omega_{i}\neq 0 and therefore the expansion (212) captures information also about the JiJ_{i}-dependence of ρ\rho. For simplicity, we will focus on line defects (p=1p=1). We will assume that d4d\geq 4 (so that m=2m=2) and discuss the modifications in the case d=3d=3 in the end of this section.

Using (211) we obtain the specialised version of (212)

𝑑Δ𝑑J1𝑑J2ρ(Δ,J1,J2)rΔeiJ1θ1iJ2θ2=exp{Seff(0)+O(Lp+1)}\displaystyle\int d\Delta dJ_{1}dJ_{2}\rho(\Delta,J_{1},J_{2})r^{\Delta}e^{-iJ_{1}\theta_{1}-iJ_{2}\theta_{2}}=\exp\left\{-S_{\text{eff}}^{(0)}+O(L^{-p+1})\right\}
=exp{4a0L1+Ω22K(Ω12+Ω221+Ω22)+O(Lp+1)},\displaystyle\quad=\exp\left\{\frac{4a_{0}}{L\sqrt{1+\Omega_{2}^{2}}}K\left(\frac{-\Omega_{1}^{2}+\Omega_{2}^{2}}{1+\Omega_{2}^{2}}\right)+O(L^{-p+1})\right\}, (221)

where r,θir,\theta_{i} are related to L,ΩiL,\Omega_{i} by (208). The density ρ(Δ,J1,J2)\rho(\Delta,J_{1},J_{2}) can be now obtained using inverse Fourier-Laplace transform

ρ(Δ,J1,J2)=1(2π)3i𝑑θ1𝑑θ2y0iy0+i𝑑yexp{Seff(0)+O(yp+1)}eΔy+iJ1θ1+iJ2θ2.\displaystyle\rho(\Delta,J_{1},J_{2})=\frac{1}{(2\pi)^{3}i}\int d\theta_{1}d\theta_{2}\int_{y_{0}-i\infty}^{y_{0}+i\infty}dy\exp\left\{-S^{(0)}_{\text{eff}}+O(y^{-p+1})\right\}e^{\Delta y+iJ_{1}\theta_{1}+iJ_{2}\theta_{2}}. (222)

We can again obtain a prediction for ρ(Δ,J1,J2)\rho(\Delta,J_{1},J_{2}) at large Δ\Delta under the assumption that the above integrals can be computed using a saddle point approximation with a saddle at small yy and θi\theta_{i}.

Introducing the function

F(Ω1,Ω2)=11+Ω22K(Ω12+Ω221+Ω22),\displaystyle F(\Omega_{1},\Omega_{2})=\frac{1}{\sqrt{1+\Omega_{2}^{2}}}K\left(\frac{-\Omega_{1}^{2}+\Omega_{2}^{2}}{1+\Omega_{2}^{2}}\right), (223)

the saddle-point equations become (recall θi=LΩiyΩi\theta_{i}=L\Omega_{i}\approx y\Omega_{i})

4a0y2F(Ω1,Ω2)+Δ+iJ1Ω1+iJ2Ω2=0,4a0yF(Ω1,Ω2)Ωi+iyJi=0.\displaystyle-\frac{4a_{0}}{y^{2}}F(\Omega_{1},\Omega_{2})+\Delta+iJ_{1}\Omega_{1}+iJ_{2}\Omega_{2}=0,\quad\frac{4a_{0}}{y}\frac{\partial F(\Omega_{1},\Omega_{2})}{\partial\Omega_{i}}+iyJ_{i}=0. (224)

In other words,

y=4a0F(Ω1,Ω2)Δ+iJ1Ω1+iJ2Ω2,\displaystyle\qquad\quad y=\sqrt{\frac{4a_{0}F(\Omega_{1},\Omega_{2})}{\Delta+iJ_{1}\Omega_{1}+iJ_{2}\Omega_{2}}}, (225)
logF(Ω1,Ω2)Ωi=iJiΔ+iJ1Ω1+iJ2Ω2.\displaystyle\frac{\partial\log F(\Omega_{1},\Omega_{2})}{\partial\Omega_{i}}=-\frac{iJ_{i}}{\Delta+iJ_{1}\Omega_{1}+iJ_{2}\Omega_{2}}. (226)

The leading one-point function density is

ρ(Δ,J1,J2)exp[4a0(Δ+iJ1Ω1+iJ2Ω2)F(Ω1,Ω2)],\displaystyle\rho(\Delta,J_{1},J_{2})\sim\exp\left[4\sqrt{a_{0}(\Delta+iJ_{1}\Omega_{1}+iJ_{2}\Omega_{2})F(\Omega_{1},\Omega_{2})}\right], (227)

where Ω1\Omega_{1} and Ω2\Omega_{2} are determined from (226) with FF defined in (223). The solution for Ωi\Omega_{i} is purely imaginary, and for JiΔJ_{i}\ll\Delta can be expanded as

iΩ1=2j13j1(j1j2)(j1+j2)+32j1(j1j2)(j1+j2)(5j12j22)+O(j1,26),\displaystyle-i\Omega_{1}=2j_{1}-3j_{1}(j_{1}-j_{2})(j_{1}+j_{2})+\frac{3}{2}j_{1}(j_{1}-j_{2})(j_{1}+j_{2})(5j_{1}^{2}-j_{2}^{2})+O(j^{6}_{1,2}), (228)

where ji=Ji/Δj_{i}=J_{i}/\Delta. The expression for Ω2\Omega_{2} is obtained by swapping 11 and 22 above. The corresponding expansion for the one-point function density is

logρ(Δ,J1,J2)\displaystyle\log\rho(\Delta,J_{1},J_{2})
8πa0Δ(112(j12+j22)32j12j2214(j12+j22)(j14+5j12j22+j24)+O(j1,28)).\displaystyle\quad\sim\sqrt{8\pi a_{0}\Delta}\left(1-\tfrac{1}{2}(j_{1}^{2}+j_{2}^{2})-\tfrac{3}{2}j_{1}^{2}j_{2}^{2}-\tfrac{1}{4}(j_{1}^{2}+j_{2}^{2})(j_{1}^{4}+5j_{1}^{2}j_{2}^{2}+j_{2}^{4})+O(j_{1,2}^{8})\right). (229)

The above discussion applies to line defects in d4d\geq 4. The case d=3d=3 can be obtained by setting J2=Ω2=0J_{2}=\Omega_{2}=0. In particular, Ω=Ω1\Omega=\Omega_{1} is determined by J=J1J=J_{1} via the specialised version of (226)

logK(Ω2)Ω=iJΔ+iJΩ.\displaystyle\frac{\partial\log K(-\Omega^{2})}{\partial\Omega}=-\frac{iJ}{\Delta+iJ\Omega}. (230)

In this case it is easy to check that the solution Ω(j)\Omega(j) maps j[1,1]j\in[-1,1] to iΩ[1,1]-i\Omega\in[-1,1] monotonically, with Ω(±1)=±i\Omega(\pm 1)=\pm i. The leading one-point function density is

ρ(Δ,J)exp[4a0(Δ+iJΩ)K(Ω2)].\displaystyle\rho(\Delta,J)\sim\exp\left[4\sqrt{a_{0}(\Delta+iJ\Omega)K(-\Omega^{2})}\right]. (231)

The expansion in small j=J/Δj=J/\Delta is obtained by setting j1=j,j2=0j_{1}=j,j_{2}=0 in (5.4),

logρ(Δ,J)8πa0Δ(112j214j6+O(j1,28)).\displaystyle\log\rho(\Delta,J)\sim\sqrt{8\pi a_{0}\Delta}\left(1-\tfrac{1}{2}j^{2}-\tfrac{1}{4}j^{6}+O(j_{1,2}^{8})\right). (232)

6 Anomalies

In this section we discuss how presence of Weyl anomalies affects the construction of the effective action SeffS_{\text{eff}}. We show that methods of section 2 allow a general construction of Weyl-anomaly matching terms.

Note that conformal defects might also have associated gravitational or other anomalies. To simplify the discussion, we assume that there no anomalies are present other than Weyl anomalies. In general, matching of all anomalies has to be carefully taken into account.

6.1 Weyl anomalies

Generally speaking, the partition functions of a conformal field theory are not necessarily conformally-invariant, but can transform anomalously under Weyl transformation,

𝒵(e2ωg)=e𝒜(g,ω)𝒵(g).\displaystyle\mathcal{Z}(e^{2\omega}g)=e^{\mathcal{A}(g,\omega)}\mathcal{Z}(g). (233)

Here 𝒵(g)\mathcal{Z}(g) denotes the partition function (possibly with local operators and defects inserted), and 𝒜(g,ω)\mathcal{A}(g,\omega) is the Weyl anomaly. The Weyl anomaly has to be a local functional, and is usually specified to the leading order in ω\omega. For example, in a 4d bulk CFT, with no operator insertions in the partition function,

𝒜(g,ω)=d4xgω(aE4+cCμνραCμνρα+ic~εμνραRμνγβRγβ)ρα+O(ω2)\displaystyle\mathcal{A}(g,\omega)=\int d^{4}x\sqrt{g}\,\omega\,(-aE_{4}+cC_{\mu\nu\rho\alpha}C^{\mu\nu\rho\alpha}+i\widetilde{c}\varepsilon^{\mu\nu\rho\alpha}R_{\mu\nu\gamma\beta}R^{\gamma\beta}{}_{\rho\alpha})+O(\omega^{2}) (234)

for some anomaly coefficients a,c,c~a,c,\widetilde{c} (see Duff:1993wm ). When defects are inserted into the partition function, the anomaly splits into the bulk and defect contributions. For example, if we have defects 𝒟1𝒟n\mathcal{D}_{1}\cdots\mathcal{D}_{n} inserted, then

𝒜(g,ω)=𝒜bulk(g,ω)+i=1n𝒜𝒟i(g,ω),\displaystyle\mathcal{A}(g,\omega)=\mathcal{A}_{\text{bulk}}(g,\omega)+\sum_{i=1}^{n}\mathcal{A}_{\mathcal{D}_{i}}(g,\omega), (235)

where each 𝒜𝒟i\mathcal{A}_{\mathcal{D}_{i}} is a local functional on the defect 𝒟i\mathcal{D}_{i}. See Chalabi:2021jud for a review of defect Weyl anomalies.

In this section we will address two problems. The first problem is that in the fusion identity (2)

𝒟1𝒟2𝒟Σ[eSeff]\displaystyle\mathcal{D}_{1}\mathcal{D}_{2}\sim\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}}] (236)

we have different defects on the two sides, and therefore different Weyl anomalies. If SeffS_{\text{eff}} is Weyl-invariant, the two sides will not in general transform in the same way under Weyl transformations. This means that SeffS_{\text{eff}} needs to be modified in order to match the Weyl anomalies. We explicitly construct the necessary anomaly-matching terms in section 6.2.

The second problem is that, as discussed in section 3.8, we in general need to consider transverse-symmetry breaking defects 𝒟Σ\mathcal{D}_{\Sigma}. We are not aware of a classification of Weyl anomalies for such defects, and therefore in sections 6.3 and 6.4 we classify Weyl anomalies for line and surface defects that break transverse rotations to SO(q1)\mathrm{SO}(q-1).

6.2 Anomaly-matching

First, note that under a Weyl transformation, a partition function involving the left-hand side of (236) transforms as

𝒵(e2ωg,𝒟1𝒟2)=e𝒜bulk(g,ω)+𝒜𝒟1(g,ω)+𝒜𝒟2(g,ω)𝒵(g,𝒟1𝒟2).\displaystyle\mathcal{Z}(e^{2\omega}g,\mathcal{D}_{1}\mathcal{D}_{2})=e^{\mathcal{A}_{\text{bulk}}(g,\omega)+\mathcal{A}_{\mathcal{D}_{1}}(g,\omega)+\mathcal{A}_{\mathcal{D}_{2}}(g,\omega)}\mathcal{Z}(g,\mathcal{D}_{1}\mathcal{D}_{2}). (237)

On the other hand, a partition function with the right-hand side transforms as

𝒵(e2ωg,𝒟Σ[eSeff(e2ωg)])=e𝒜bulk(g,ω)+𝒜𝒟Σ(g,ω)+Seff(g)Seff(e2ωg)𝒵(g,𝒟Σ[eSeff(g)]).\displaystyle\mathcal{Z}(e^{2\omega}g,\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}(e^{2\omega}g)}])=e^{\mathcal{A}_{\text{bulk}}(g,\omega)+\mathcal{A}_{\mathcal{D}_{\Sigma}}(g,\omega)+S_{\text{eff}}(g)-S_{\text{eff}}(e^{2\omega}g)}\mathcal{Z}(g,\mathcal{D}_{\Sigma}[e^{-S_{\text{eff}}(g)}]). (238)

Here, we explicitly keep the dependence of the effective action on the metric gg and we also assumed that Seff(g)Seff(e2ωg)S_{\text{eff}}(g)-S_{\text{eff}}(e^{2\omega}g) only contains the identity operator contribution and thus can be taken out of the partition function. An important comment is in order regarding 𝒜𝒟Σ(g,ω)\mathcal{A}_{\mathcal{D}_{\Sigma}}(g,\omega). This anomaly is to be computed for 𝒟Σ\mathcal{D}_{\Sigma}, displaced according to the displacement operator coupling as discussed in section 3.7.474747There are two ways of seeing this. On the one hand, adding a displacement operator coupling is by definition equivalent to displacing the defect. Therefore, the partition function with the displacement operator coupling turned on and the partition function with the displaced 𝒟Σ\mathcal{D}_{\Sigma} should behave in the same way under Weyl transformations. On the other hand, if we use conformal perturbation theory in the displacement coupling, Weyl anomaly contributions will come from divergences associated with the identity operator appearing in repeated OPEs of DμD_{\mu}. Consistency of these two points of view leads to relations between correlation functions of DμD_{\mu} and defect Weyl anomaly coefficients Herzog:2017xha ; Herzog:2017kkj . Other subtleties in the evaluation of 𝒜𝒟Σ(g,ω)\mathcal{A}_{\mathcal{D}_{\Sigma}}(g,\omega) might arise if SeffS_{\text{eff}} contains operators with special scaling dimensions, see section 3.7 for a discussion of this.

The Weyl anomaly will be matched if

𝒜𝒟1(g,ω)+𝒜𝒟2(g,ω)𝒜𝒟Σ(g,ω)=Seff(g)Seff(e2ωg).\displaystyle\mathcal{A}_{\mathcal{D}_{1}}(g,\omega)+\mathcal{A}_{\mathcal{D}_{2}}(g,\omega)-\mathcal{A}_{\mathcal{D}_{\Sigma}}(g,\omega)=S_{\text{eff}}(g)-S_{\text{eff}}(e^{2\omega}g). (239)

To proceed, we note that the transformation law (233) implies the following addition law for 𝒜(g,ω)\mathcal{A}(g,\omega),

𝒜(e2ω2g,ω1)=𝒜(g,ω1+ω2)𝒜(g,ω2).\displaystyle\mathcal{A}(e^{2\omega_{2}}g,\omega_{1})=\mathcal{A}(g,\omega_{1}+\omega_{2})-\mathcal{A}(g,\omega_{2}). (240)

Consider now the term

(g)=𝒜𝒟1(g,log(g)),\displaystyle\mathcal{B}(g)=\mathcal{A}_{\mathcal{D}_{1}}(g,-\log\ell(g)), (241)

where we keep explicitly the dependence of \ell on gg. From (240), this transforms as

(e2ωg)=𝒜𝒟1(e2ωg,ωlog(g))=(g)𝒜𝒟1(g,ω).\displaystyle\mathcal{B}(e^{2\omega}g)=\mathcal{A}_{\mathcal{D}_{1}}(e^{2\omega}g,-\omega-\log\ell(g))=\mathcal{B}(g)-\mathcal{A}_{\mathcal{D}_{1}}(g,\omega). (242)

Therefore, if we write

Seff=𝒜𝒟1(g,log)+𝒜𝒟2(g,log)𝒜𝒟Σ(g,log)+Seff0,\displaystyle S_{\text{eff}}=\mathcal{A}_{\mathcal{D}_{1}}(g,-\log\ell)+\mathcal{A}_{\mathcal{D}_{2}}(g,-\log\ell)-\mathcal{A}_{\mathcal{D}_{\Sigma}}(g,-\log\ell)+S^{0}_{\text{eff}}, (243)

then (239) implies that Seff0(g)S^{0}_{\text{eff}}(g) is Weyl-invariant and we can use the discussion of the previous sections to construct it. The terms

𝒜𝒟1(g,log)+𝒜𝒟2(g,log)𝒜𝒟Σ(g,log)\displaystyle\mathcal{A}_{\mathcal{D}_{1}}(g,-\log\ell)+\mathcal{A}_{\mathcal{D}_{2}}(g,-\log\ell)-\mathcal{A}_{\mathcal{D}_{\Sigma}}(g,-\log\ell) (244)

are therefore the necessary Weyl-anomaly matching terms.

Note that in (244) different terms evaluate log-\log\ell on a different submanifolds – the Weyl anomaly of 𝒟1\mathcal{D}_{1} depends on the Weyl factor on 𝒟1\mathcal{D}_{1} and so on. It is therefore crucial for this construction to use the bulk-extended version of \ell that was constructed in section 2.3. When evaluated in terms of data on 𝒟Σ\mathcal{D}_{\Sigma} (as all other terms in the effective action), (244) becomes an infinite derivative expansion, suppressed by powers of LL.

Finally, let us briefly discuss how one can reconstruct 𝒜(g,ω)\mathcal{A}(g,\omega) from the its leading term in ω\omega, which is the term that is usually specified. Suppose

𝒜(g,ω)=𝒜(1)(g,ω)+O(ω2),\displaystyle\mathcal{A}(g,\omega)=\mathcal{A}^{(1)}(g,\omega)+O(\omega^{2}), (245)

where 𝒜(1)(g,ω)\mathcal{A}^{(1)}(g,\omega) is linear in ω\omega. First, we note that for any Weyl transformation ω~\widetilde{\omega},

𝒜(1)(e2ω~g,ω)=𝒜(1)(g,ω)+δ𝒜(1)(g,ω;ω~),\displaystyle\mathcal{A}^{(1)}(e^{2\widetilde{\omega}}g,\omega)=\mathcal{A}^{(1)}(g,\omega)+\delta\mathcal{A}^{(1)}(g,\omega;\partial\widetilde{\omega}), (246)

where δ𝒜(1)(g,ω;ω~)\delta\mathcal{A}^{(1)}(g,\omega;\partial\widetilde{\omega}) is a finite-order local polynomial in ω~\widetilde{\omega}, and furthermore ω~\widetilde{\omega} only enters through its derivatives. This is because 𝒜(1)(g,ω)\mathcal{A}^{(1)}(g,\omega) is scale-invariant and δ𝒜(1)\delta\mathcal{A}^{(1)} must vanish for constant ω~\widetilde{\omega}. But then δ𝒜(1)\delta\mathcal{A}^{(1)} only depends on derivatives of ω~\widetilde{\omega}, and only finitely many derivatives can enter by dimensional analysis. For example, for a line defect at most ω~\partial\widetilde{\omega} can appear and therefore δ𝒜(1)(g,ω;ω~)\delta\mathcal{A}^{(1)}(g,\omega;\partial\widetilde{\omega}) is linear in ω~\partial\widetilde{\omega}. On the other hand, for a surface defect we can have ω~\partial\widetilde{\omega}, 2ω~\partial^{2}\widetilde{\omega}, and (ω~)2(\partial\widetilde{\omega})^{2}.

It remains to note that the composition law (240) implies that

t𝒜(g,tω)=𝒜(1)(e2tωg,ω)=𝒜(1)(g,ω)+δ𝒜(1)(g,ω;tω).\displaystyle\partial_{t}\mathcal{A}(g,t\omega)=\mathcal{A}^{(1)}(e^{2t\omega}g,\omega)=\mathcal{A}^{(1)}(g,\omega)+\delta\mathcal{A}^{(1)}(g,\omega;t\partial\omega). (247)

This implies

𝒜(g,ω)=𝒜(1)(g,ω)+01𝑑tδ𝒜(1)(g,ω;tω).\displaystyle\mathcal{A}(g,\omega)=\mathcal{A}^{(1)}(g,\omega)+\int_{0}^{1}dt\delta\mathcal{A}^{(1)}(g,\omega;t\partial\omega). (248)

Since the integrand in the right-hand side is an explicit polynomial in tt that is easily computable from 𝒜(1)\mathcal{A}^{(1)}, this expression allows one to compute 𝒜(g,ω)\mathcal{A}(g,\omega).

6.3 Line defects

In this section we classify Weyl anomalies for line defects 𝒟\mathcal{D} that break transverse rotations down to SO(q1)\mathrm{SO}(q-1). We find a non-trivial anomaly in d=3d=3 but no anomalies in d4d\geq 4.

We follow the standard strategy (see, for example, Chalabi:2021jud ), where we first the most general local ansatz for 𝒜𝒟(1)\mathcal{A}^{(1)}_{\mathcal{D}}, impose Wess-Zumino consistency conditions, and finally isolate the scheme-invariant part of the anomaly. We first consider d=3d=3 and generalise to d4d\geq 4 later.

Recall from section 3.8 that the breaking of transverse rotations is parameterised by a unit normal vector field nμn^{\mu} of scaling dimension 0. This vector field allows us to construct new terms in the ansatz for 𝒜𝒟(1)\mathcal{A}^{(1)}_{\mathcal{D}}, making the classification problem different from the case of transverse rotation-preserving defects.

For convenience, we introduce a unit vector field τ\tau tangent to the defect. Note that in this section we work with the physical metric gμνg_{\mu\nu}. Since τ\tau is a unit vector, it also has dimension 0. Using τ\tau and nn, we can construct a unit vector mm orthogonal to both of them,

mμ=εμnννρτρ,\displaystyle m^{\mu}=\varepsilon^{\mu}{}_{\nu\rho}n^{\nu}\tau^{\rho}, (249)

The most general form of the Weyl anomaly is

𝒜𝒟(1)(g,ω)=𝑑zγf,\displaystyle\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega)=\int\,dz\sqrt{\gamma}f, (250)

where ff is a local expression of mass dimension 1. We first construct the most general ansatz for ff using the standard curvature invariants, n,τ,mn,\tau,m, and covariant derivatives. Only linearly independent contributions to 𝒜𝒟(1)\mathcal{A}^{(1)}_{\mathcal{D}} should be included. In particular, in ff we should not include total derivatives. This leads to the following ansatz,

𝒜𝒟(1)(g,ω)=\displaystyle\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega)= 𝑑zγω(c1nμ\RomanbarIIμ+c2mμ\RomanbarIIμ+ic3τamμanμ)\displaystyle\int dz\sqrt{\gamma}\omega\left(c_{1}n_{\mu}\Romanbar{II}^{\mu}+c_{2}m_{\mu}\Romanbar{II}^{\mu}+ic_{3}\tau^{a}m_{\mu}\nabla^{\perp}_{a}n^{\mu}\right)
+𝑑zγμω(c4nμ+c5mμ).\displaystyle+\int dz\sqrt{\gamma}\partial_{\mu}\omega\left(c_{4}n^{\mu}+c_{5}m^{\mu}\right). (251)

The Weyl variation of 𝒜𝒟(1)(g,ω1)\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega_{1}) is given by

δω2𝒜𝒟(1)(g,ω1)=𝑑zγ(c1nμω1μω2+c2mμω1μω2).\displaystyle\delta_{\omega_{2}}\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega_{1})=-\int dz\sqrt{\gamma}\left(c_{1}n^{\mu}\omega_{1}\partial_{\mu}\omega_{2}+c_{2}m^{\mu}\omega_{1}\partial_{\mu}\omega_{2}\right). (252)

Wess-Zumino consistency condition requires δω1𝒜𝒟(1)(g,ω2)=δω2𝒜𝒟(1)(g,ω1)\delta_{\omega_{1}}\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega_{2})=\delta_{\omega_{2}}\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega_{1}), which implies c1=c2=0c_{1}=c_{2}=0. Therefore, the most general form for the anomaly allowed by Wess-Zumino consistency condition is

𝒜𝒟(1)(g,ω)=𝑑zγ(ic3ωτamμanμ+μω(c4nμ+c5mμ)).\displaystyle\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega)=\int dz\sqrt{\gamma}\left(ic_{3}\omega\tau^{a}m_{\mu}{\nabla}^{\perp}_{a}n^{\mu}+\partial_{\mu}\omega(c_{4}n^{\mu}+c_{5}m^{\mu})\right). (253)

The coefficients c4c_{4} and c5c_{5} can be set to 0 using local counter-terms n\RomanbarIIn\cdot\Romanbar{II} and m\RomanbarIIm\cdot\Romanbar{II}. Therefore, the final form of the anomaly is

𝒜𝒟(1)(g,ω)\displaystyle\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega) =i𝑑zγc3ωτamμanμ.\displaystyle=i\int dz\sqrt{\gamma}c_{3}\omega\tau^{a}m_{\mu}{\nabla}^{\perp}_{a}n^{\mu}. (254)

Note that for a constant ω\omega this reduces to 2πic3I(n)2\pi ic_{3}I(n) evaluated in the physical metric, where I(n)I(n) is discussed in section 3.6.

The only essential modification in d>3d>3 is that the terms that involve mμm^{\mu} do not exist there. The result is that no defect Weyl anomaly is possible in that case.

6.4 Surface defects in d=4d=4

We now consider surface defects in d=4d=4. We once again assume that this defect breaks rotational symmetry along the defect, and that this breaking can be parameterised by a normal vector field nμn^{\mu}. As the defect is now two-dimensional, we can choose an orthonormal basis in its tangent space at every point, parameterised by vector fields τa\tau^{a} and σa\sigma^{a}. This choice of basis is however arbitrary and unphysical, and therefore the terms of the Weyl anomaly should not depend on this choice. The easiest way to impose such a requirement, is to use complex coordinates

ζa=τa+iσa2,ζ¯a=τaiσa2\displaystyle\zeta^{a}=\frac{\tau^{a}+i\sigma^{a}}{\sqrt{2}},\qquad\overline{\zeta}^{a}=\frac{\tau^{a}-i\sigma^{a}}{\sqrt{2}} (255)

for the tangent bundle of the defect. Rotations of these basis vectors map ζaeiθζa\zeta^{a}\to e^{i\theta}\zeta^{a}, ζ¯aeiθζ¯a\overline{\zeta}^{a}\to e^{-i\theta}\overline{\zeta}^{a}, and so we can easily see how terms transform under these rotations.

As before, we also introduce a final basis vector normal to the defect,

mμ=εμnννρστρsσ,\displaystyle m^{\mu}=\varepsilon^{\mu}{}_{\nu\rho\sigma}n^{\nu}\tau^{\rho}s^{\sigma}, (256)

and we then can use

χμ=mμ+inμ2,χ¯μ=mμinμ2\displaystyle\chi^{\mu}=\frac{m^{\mu}+in^{\mu}}{\sqrt{2}},\qquad\overline{\chi}^{\mu}=\frac{m^{\mu}-in^{\mu}}{\sqrt{2}} (257)

as complex coordinates on the defect. Note that mm is invariant under rotations of τ,σ\tau,\sigma. Now in each tangent space we have a set of coordinates defined by ζ,ζ¯,χ,χ¯\zeta,\overline{\zeta},\chi,\overline{\chi}, and we can use these coordinates to describe the components of various tensors, e.g. \RomanbarIIζζχ\Romanbar{II}^{\chi}_{\zeta\zeta}.

We then proceed as in the case of the line defect. As compared with the last section, we shall give fewer details of the calculations involved in this case. This is because the methods are almost identical, but the list of terms (given in appendix D) we have to consider is much longer. This list was once again obtained by writing down a list of all possible terms with two derivatives, and removing any terms that are not linearly independent. In order to remove such linear dependence, the Gauss, Codazzi and Ricci identities have to be considered.

After applying Wess-Zumino consistency conditions taking local counterterms into account, we find that the scheme-independent part of the defect Weyl anomaly is

𝒜𝒟(1)(g,ω)=\displaystyle\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega)= d2zγω(d1χ¯μ()2χμ+d¯1χμ()2χ¯μ+d2\RomanbarII̊χζζ\RomanbarII̊χζ¯ζ¯+d¯2\RomanbarII̊χ¯ζζ\RomanbarII̊χ¯ζ¯ζ¯\displaystyle\int d^{2}z\sqrt{\gamma}\omega\bigg{(}d_{1}\overline{\chi}_{\mu}(\nabla^{\perp})^{2}\chi^{\mu}+\overline{d}_{1}\chi_{\mu}(\nabla^{\perp})^{2}\overline{\chi}^{\mu}+d_{2}\mathring{\Romanbar{II}}_{\chi\zeta\zeta}\mathring{\Romanbar{II}}_{\chi\overline{\zeta}\overline{\zeta}}+\overline{d}_{2}\mathring{\Romanbar{II}}_{\overline{\chi}\zeta\zeta}\mathring{\Romanbar{II}}_{\overline{\chi}\overline{\zeta}\overline{\zeta}}
+e1\RomanbarII̊χζζ\RomanbarII̊χ¯ζ¯ζ¯+e¯1\RomanbarII̊χ¯ζζ\RomanbarII̊χζ¯ζ¯+e2Cχχ¯χχ¯+ie3Cχχ¯ζζ¯+c𝒟Rγ)\displaystyle\hskip 62.59596pt+e_{1}\mathring{\Romanbar{II}}_{\chi\zeta\zeta}\mathring{\Romanbar{II}}_{\overline{\chi}\overline{\zeta}\overline{\zeta}}+\overline{e}_{1}\mathring{\Romanbar{II}}_{\overline{\chi}\zeta\zeta}\mathring{\Romanbar{II}}_{\chi\overline{\zeta}\overline{\zeta}}+e_{2}C_{\chi\overline{\chi}\chi\overline{\chi}}+ie_{3}C_{\chi\overline{\chi}\zeta\overline{\zeta}}+c_{\mathcal{D}}R_{\gamma}\bigg{)}
+id2zγd5(\RomanbarIIχχω\RomanbarIIχ¯χ¯ω),\displaystyle+i\int d^{2}z\sqrt{\gamma}d_{5}(\Romanbar{II}_{\chi}\partial_{\chi}\omega-\Romanbar{II}_{\overline{\chi}}\partial_{\overline{\chi}}\omega), (258)

where RγR_{\gamma} is the Ricci scalar of the defect metric. The coefficients e2,e3,c𝒟,d5e_{2},e_{3},c_{\mathcal{D}},d_{5}, are real.484848Note that complex conjugation plus orientation reversal exchanges ww and w¯\overline{w} and does nothing to ζ\zeta and ζ¯\overline{\zeta}. The terms with coefficients c𝒟c_{\mathcal{D}} and eie_{i} do not break transverse rotations and are already known Chalabi:2021jud ; Graham:1999pm ; Schwimmer:2008yh ; Cvitan:2015ana ; Jensen:2018rxu , whereas the terms with coefficients did_{i} do break transverse rotations, and are new. One point to note is that the coefficients did_{i} and eie_{i} can generically be position-dependent, but by Wess-Zumino consistency conditions, c𝒟c_{\mathcal{D}} must be position-independent and thus independent of marginal couplings. All of the terms are B-type (i.e. they transform trivially under Weyl transformations) except from c𝒟Rγc_{\mathcal{D}}R_{\gamma}, which is A-type.

7 Fusion and OPE

In this section we discuss the analogies and differences between the fusion of conformal defects and the OPE of local operators. Formally, a local operator can be viewed as a 0-dimensional defect, which allows many questions to be translated between the two situations. As everywhere in the paper, in this section “conformal defect” refers to a conformal defect with p>0p>0, unless explicitly stated otherwise.

We first discuss the differences in how one labels conformal defects and primary operators. To a primary local operator 𝒪{\mathcal{O}} one associates a set of quantum numbers, which include its scaling dimension, spin representation, and representations under various global symmetry groups. By contrast, no such assignment is made in the case of conformal defects.

Consider first the spin and global symmetry representations. These describe the action of various symmetry groups on the vector space spanned by different components of the local operator 𝒪{\mathcal{O}}. In other words, a given component of 𝒪{\mathcal{O}} transforms non-trivially under these symmetries. This of course can also happen for conformal defects, which can break global symmetries and transverse rotations. A crucial difference is, however, that the local operators form a vector space, and therefore organise into a linear representation of the symmetry. This representation can then be decomposed into irreducible components, which is what allows us to label local operators by irreps. By contrast, as discussed in section 3.5, locality requirements mean that conformal defects do not form a natural vector space, and therefore no such decomposition is possible.494949In this sense the “spinning conformal defects” of Guha:2018snh ; Kobayashi:2018okw are necessarily non-local and are not the ones considered in this paper.

The scaling dimensions of local operators also fall under the above discussion since they determine the representation of the dilatation subgroup. However, in this case a different point of view is possible. Specifically, one can interpret the scaling dimension of a local operator as the Weyl anomaly for a 0-dimensional conformal defect. For higher-dimensional defects the classification of Weyl anomalies is different Chalabi:2021jud , and it is therefore not surprising that there isn’t a direct analogue of scaling dimensions for all conformal defects.

From this point of view, the leading-order term in the local operator product expansion can be seen as a Weyl anomaly-matching term in the effective action. For example, consider the contribution of a scalar local operator 𝒪{\mathcal{O}} of dimension Δ\Delta to the OPE of two scalar operators ϕ\phi of dimension Δϕ\Delta_{\phi},

ϕ(x)ϕ(0)|x|Δ2Δϕ𝒪(0)+=e(Δ2Δϕ)log|x|𝒪(0)+.\displaystyle\phi(x)\phi(0)\ni|x|^{\Delta-2\Delta_{\phi}}{\mathcal{O}}(0)+\cdots=e^{(\Delta-2\Delta_{\phi})\log|x|}{\mathcal{O}}(0)+\cdots. (259)

This can be written using a notation similar to the one we used for conformal defects,

ϕ(v)ϕ(0)=e(Δ2Δϕ)log𝒪(0)+.\displaystyle\phi(v)\phi(0)=e^{(\Delta-2\Delta_{\phi})\log\ell}{\mathcal{O}}(0)+\cdots. (260)

Here, we parameterise the position of one of the operators using a tangent vector vv at 0, and 2=gμνvμvν\ell^{2}=g_{\mu\nu}v^{\mu}v^{\nu}. Since on a 0-dimensional defect integration is trivial, this can be viewed the leading-order analogue of equation (244) for conformal defects. From section 6.2 we know that this leading-order term fails to completely match the anomaly, since ϕ(v)\phi(v) and ϕ(0)\phi(0) transform with the Weyl factor evaluated at different points.

In fact, it impossible to complete (260) into an equation that matches the Weyl anomaly to all orders in LL without including some analogue of non-trivial operators in SeffS_{\text{eff}}. Indeed, if it were possible, the OPE ϕ(v)ϕ(0)\phi(v)\phi(0) would be conformally-invariant even with only the primary 𝒪(0){\mathcal{O}}(0) is included, but it is well-known that inclusion of descendants is generally necessary. The anomaly-matching terms constructed in section 6 do not exist here because the differential equation (19) and the invariant (67) are not defined for p=0p=0.505050For the same reason, the conformally-invariant definition of vv in section 2.1 does not work for 0-dimensional defects. In fact it is not hard to show that no conformally-invariant definition of displacement vector vv is possible for p=0p=0: the group of conformal transformations that preserve two points in d\mathbb{R}^{d} is too large – it can rotate and rescale any tangent vector vv without moving the insertion points of the local operators. Therefore, no tangent vector can parameterise the relative position of two points in a conformally-invariant way.

To fix this problem, we should include the descendants of 𝒪{\mathcal{O}}. It is natural to treat these descendants as being analogous to insertions of a displacement operator on a 0-dimensional defect 𝒟\mathcal{D},

μ1μn𝒪𝒟[Dμ1Dμn].\displaystyle\langle\partial_{\mu_{1}}\cdots\partial_{\mu_{n}}{\mathcal{O}}\cdots\rangle\sim\langle\mathcal{D}[D_{\mu_{1}}\cdots D_{\mu_{n}}]\cdots\rangle. (261)

The nn-point correlation function of DμD_{\mu} are subject to scheme ambiguities which affect its definition at coincident points (c.f. section 3.7). In a generic situation, the possible modifications are given by terms including derivatives of delta-functions and lower-point correlation functions of DμD_{\mu}. In the case of 0-dimensional defects, the only possible configuration of nn points is a coincident point configuration, which in terms of 𝒪{\mathcal{O}} means that we can correct the nnth order derivatives of 𝒪{\mathcal{O}} by its lower-order derivatives and interpret this as a scheme change. For instance, we have the schemes

𝒟[Dμ1Dμn]\displaystyle\langle\mathcal{D}[D_{\mu_{1}}\cdots D_{\mu_{n}}]\cdots\rangle =μ1μn𝒪,\displaystyle=\langle\partial_{\mu_{1}}\cdots\partial_{\mu_{n}}{\mathcal{O}}\cdots\rangle, (262)
𝒟[Dμ1Dμn]\displaystyle\langle\mathcal{D}[D_{\mu_{1}}\cdots D_{\mu_{n}}]\cdots\rangle =μ1μn𝒪.\displaystyle=\langle\nabla_{\mu_{1}}\cdots\nabla_{\mu_{n}}{\mathcal{O}}\cdots\rangle. (263)

They differ by subleading derivatives of 𝒪{\mathcal{O}} times derivatives of the metric, and only the latter scheme is generally-covariant. As mentioned in footnote 50, no Weyl-invariant schemes exist in this case.

For conformal defects, we would normally include the displacement operator into the effective action as

𝒟[edpzγλμDμ],\displaystyle\mathcal{D}[e^{-\int d^{p}z\sqrt{\gamma}\lambda^{\mu}D_{\mu}}], (264)

where locality requires us to use the exponential of a local integral. This is necessary in order to obtain proper factorisation properties for path integrals. In the case of 0-dimensional defects no such requirement exists (a 0-dimensional defect cannot be cut in half) and we can therefore modify the partition functions in a more general way,

𝒟[1+λ1μDμ+12λ2μνDμDν+]𝒪+λ1μμ𝒪+12λ2μνμν𝒪+,\displaystyle\mathcal{D}[1+\lambda^{\mu}_{1}D_{\mu}+\tfrac{1}{2}\lambda^{\mu\nu}_{2}D_{\mu}D_{\nu}+\cdots]\sim{\mathcal{O}}+\lambda^{\mu}_{1}\nabla_{\mu}{\mathcal{O}}+\tfrac{1}{2}\lambda^{\mu\nu}_{2}\nabla_{\mu}\nabla_{\nu}{\mathcal{O}}+\cdots, (265)

where the coefficients λ1,λ2,\lambda_{1},\lambda_{2},\cdots are all independent. They are constrained by Weyl invariance (which in flat space fixes them completely), but not by a requirement to form an exponential series.

To summarise, we find that there are two key differences between the fusion of local operators and of general conformal defects. Firstly, the conformal geometry of a pair of conformal defects is, in a sense, simpler than that for a pair of local operators, in the sense that the Weyl-invariant parameterisation of section 2.1 and the scale function of section 2.3 exist. Due to this, Weyl-invariance appears to impose fewer constraints on fusion of conformal defects than on local operators. Secondly, the locality constraints are very different in the two cases, giving more constraints in the case of conformal defects. These two effects imply that the two operations have quite different properties despite naively being very similar.

8 Discussion

In this paper we studied effective field theory for fusion of conformal defects. We have described how such an effective field theory can be written down to any order in derivative expansion, fully taking into account the diffeomorphism- and Weyl-invariance constraints. We have also explained how the terms in the effective action can be evaluated for setups with simple kinematics. Finally, we have studied simple applications of this effective field theory: to the cusp anomalous dimensions, and to the asymptotics of bulk one-point functions.

As we explained, it is possible that transverse-symmetry breaking conformal defects might arise as the result of the fusion. We found in section 6 that such line defects in d=3d=3 can have non-trivial Weyl anomalies, and that new Weyl anomaly terms arise for surface defects in d=4d=4. It would be interesting to find and study examples of such symmetry-breaking defects.

Furthermore, it would be interesting to extend the study of cusp anomalous dimensions from section 4. For example, it is important to determine whether any non-identity operators contribute to the planar Γcusp\Gamma_{\text{cusp}} and if yes, how is it compatible with (179). Computation of subleading Wilson coefficients in the effective action using integrability techniques is another open direction.

The geometric methods that we developed in section 2 allow an easy extension to more exotic fusion setups. For example, one can consider fusion of defects of different dimensions. Alternatively, defects of the same dimension can approach each other in more general configurations, so that in the limit the have a lower-dimensional intersection. We believe these questions can be straightforwardly addressed using our approach from section 2. On the other hand, limits such as the small-radius limit of a cylindrical surface defect would require development of new tools.

Finally, an important question is to understand whether in the setup of section 5 defect fusion can be replaced by a convergent expansion of some sort. Doing so would be important for enabling the application of numerical bootstrap techniques to conformal defects. Fusion effective theory gives only an asymptotic expansion. While being sufficient for the derivation of asymptotic relations of section 5, it falls short of providing a quantitative handle on the low-energy part of the spectrum in the crossed channel.

Acknowledgments

We would like to thank Sean Curry, Nadav Drukker, Nikolay Gromov, Christopher Herzog, Zohar Komargodski and David Simmons-Duffin for discussions. We especially thank Nikolay Gromov for many helpful comments on the draft and for providing us with numerical data. The work of PK was supported by the UK Engineering and Physical Sciences Research council grant number EP/X042618/1; and the Science Technology & Facilities council grant number ST/X000753/1. RS is supported by the Royal Society-Newton International Fellowship NIF/R1/221054-Royal Society.

Appendix A Geometry and conventions

Given an embedding X:NMX:N\to M of a defect, we can construct a map TNTMTN\to TM as eaμaXμe_{a}^{\mu}\equiv\partial_{a}X^{\mu}. We shall use \nabla as notation for the Levi-Civita connections on both MM and NN, and we shall use \nabla^{\perp} for the connection in the normal bundle on NN.

When there is no risk of ambiguity we shall often commit an abuse of notation, and give defect tensors bulk indices using eaμe^{\mu}_{a}, for example defining tμeaμtat^{\mu}\equiv e^{\mu}_{a}t^{a}. We can additionally define the second fundamental form of the defect \RomanbarIIabμaebμ\Romanbar{II}^{\mu}_{ab}\equiv\nabla_{a}e^{\mu}_{b}, and as this is symmetric in aa and bb, it can be split into a trace part \RomanbarIIμ\RomanbarIIabμγab\Romanbar{II}^{\mu}\equiv\Romanbar{II}^{\mu}_{ab}\gamma^{ab}, and a traceless symmetric part \RomanbarII̊abμ\RomanbarIIabμ1p\RomanbarIIμγab\mathring{\Romanbar{II}}^{\mu}_{ab}\equiv\Romanbar{II}^{\mu}_{ab}-\frac{1}{p}\Romanbar{II}^{\mu}\gamma_{ab}.

As always, given the Levi-Civita connection on either MM or NN, we can define a Riemann tensor RρσμνR^{\rho}{}_{\sigma\mu\nu}, a Ricci tensor RμνR_{\mu\nu}, and a Ricci scalar RR. We shall additionally need the bulk Schouten tensor

Pμν1d2(Rμν12(d1)Rgμν),\displaystyle P_{\mu\nu}\equiv\frac{1}{d-2}\left(R_{\mu\nu}-\frac{1}{2(d-1)}Rg_{\mu\nu}\right), (266)

the bulk Weyl tensor

CμνρσRμνρσPμρgνσ+PνρgμσPνσgμρ+Pμσgνρ.\displaystyle C_{\mu\nu\rho\sigma}\equiv R_{\mu\nu\rho\sigma}-P_{\mu\rho}g_{\nu\sigma}+P_{\nu\rho}g_{\mu\sigma}-P_{\nu\sigma}g_{\mu\rho}+P_{\mu\sigma}g_{\nu\rho}. (267)

We define the Weyl weight nn of a quantity AA by the requirement that under a Weyl rescaling gμνe2ωgμνg_{\mu\nu}\to e^{2\omega}g_{\mu\nu} the quantity AA transforms as AenωAA\to e^{n\omega}A. Note that if AA is a tensor object with ii contravariant and jj covariant indices, A=Aν1ν2νjμ1μ2μiA=A^{\mu_{1}\mu_{2}\dots\mu_{i}}_{\nu_{1}\nu_{2}\dots\nu_{j}}, then the relation between its Weyl weight nn and what is usually meant by the scaling dimension Δ\Delta is

Δ=n+ji.\displaystyle\Delta=-n+j-i. (268)

The shift between Δ\Delta and n-n appears because the scaling dimension determines the behaviour under conformal transformations, which are coordinate transformations and act on the tensor indices. For example, the metric gμνg_{\mu\nu} has Weyl weight n=2n=2, but its scaling dimension is Δ=2+2=0\Delta=-2+2=0. It is often advantageous to work with scaling dimensions because they remain invariant when the indices are raised, lowered or contracted. Relatedly, the primary operators transform under a Weyl rescaling gμνe2ωgμνg_{\mu\nu}\to e^{2\omega}g_{\mu\nu} as

δω𝒪ν1ν2νjμ1μ2μi(x)=(Δ+ji)ω(x)𝒪ν1ν2νjμ1μ2μi(x).\displaystyle\delta_{\omega}\mathcal{O}^{\mu_{1}\mu_{2}\dots\mu_{i}}_{\nu_{1}\nu_{2}\dots\nu_{j}}(x)=(-\Delta+j-i)\omega(x)\mathcal{O}^{\mu_{1}\mu_{2}\dots\mu_{i}}_{\nu_{1}\nu_{2}\dots\nu_{j}}(x). (269)

A.1 Relations between curvature invariants

When one considers a submanifold of a Riemannian manifold, in addition to the bulk Riemann tensor RμνσρR_{\mu\nu}{}^{\rho}{}_{\sigma} it is also possible to define the intrinsic submanifold Riemann tensor Rˇabdc\check{R}_{ab}{}^{c}{}_{d} as well as the curvature tensor for the normal connection RabνμR^{\perp}_{ab}{}^{\mu}{}_{\nu}. Together with the second fundamental form \RomanbarIIabμ\Romanbar{II}^{\mu}_{ab}, these objects are not independent and are related by the Gauss, Codazzi-Mainardi, and Ricci equations. These equations (and their derivatives) are the only non-trivial equations between these curvature invariants. See SpivakIV for details.

The Gauss equation states

Rabdc\displaystyle R_{ab}{}^{c}{}_{d} =Rˇabcd2\RomanbarIIμ[a\RomanbarcIIb]dμ.\displaystyle=\check{R}_{ab}{}^{c}{}_{d}-2\Romanbar{II}_{\mu[a}{}^{c}\Romanbar{II}^{\mu}_{b]d}. (270)

The Codazzi-Mainardi equation is

Rab=μca\RomanbarIIbcμb\RomanbarIIacμ.(μ normal)\displaystyle R_{ab}{}^{\mu}{}_{c}=\nabla_{a}\Romanbar{II}^{\mu}_{bc}-\nabla_{b}\Romanbar{II}^{\mu}_{ac}.\quad(\mu\text{ normal}) (271)

Finally, the Ricci equation reads

Rab=ρσRabρσ\RomanbarIIacμ\RomanbarIIνb+c\RomanbarIIbcμ\RomanbarIIνa.c(ρ,σ normal)\displaystyle R_{ab}{}^{\rho}{}_{\sigma}=R_{ab}^{\perp}{}^{\rho}{}_{\sigma}-\Romanbar{II}^{\mu}_{ac}\Romanbar{II}_{\nu b}{}^{c}+\Romanbar{II}^{\mu}_{bc}\Romanbar{II}_{\nu a}{}^{c}.\quad(\rho,\sigma\text{ normal}) (272)

The Ricci equation (272) essentially means that we do not have to consider RR^{\perp} as a separate invariant since it can be expressed in terms of RR and \RomanbarII\Romanbar{II}. The Codazzi-Mainardi equation (271) are mostly irrelevant for this paper since we do not ever have to consider covariant derivatives of \RomanbarII\Romanbar{II}. On the other hand, the Gauss equation (270) does reduce the number of independent two-derivative terms in certain situations.

It is only the fully-contracted Gauss equation that can appear in our analysis. This gives

Rabab\displaystyle R_{ab}{}^{ab} =Rˇ\RomanbarIIμ\RomanbarIIμ+\RomanbarIIμb\RomanbaraIIaμ.b\displaystyle=\check{R}-\Romanbar{II}_{\mu}\Romanbar{II}^{\mu}+\Romanbar{II}_{\mu b}{}^{a}\Romanbar{II}^{\mu}_{a}{}^{b}. (273)

Notice that the left-hand side has the bulk Riemann tensor contracted only along the defect directions. Rearranging the sums, we can rewrite this as Chalabi:2021jud

R\displaystyle R =Rˇ\RomanbarIIμ\RomanbarIIμ+\RomanbarIIμab\RomanbarIIμab+2RjjRjk,jk\displaystyle=\check{R}-\Romanbar{II}_{\mu}\Romanbar{II}^{\mu}+\Romanbar{II}_{\mu ab}\Romanbar{II}^{\mu ab}+2R^{j}_{j}-R_{jk}{}^{jk}, (274)

where the contractions over j,kj,k are taken in the normal directions.

We will also need the following special case of this identity. Assuming that normal components of Schouten tensor vanish as in (84), and that \RomanbarII^μ=0\widehat{\Romanbar{II}}^{\mu}=0 as in (82) we then find in the fusion metric (using also (267))

d1qd1R^\displaystyle\frac{d-1-q}{d-1}\widehat{R} =Rˇ^+\RomanbarII^μab\RomanbarII^μabCjk.jk\displaystyle=\widehat{\check{R}}+\widehat{\Romanbar{II}}_{\mu ab}\widehat{\Romanbar{II}}^{\mu ab}-C_{jk}{}^{jk}. (275)

For codimension q=1q=1 the contraction of the Weyl tensor vanishes by anti-symmetry and we get

d2d1R^\displaystyle\frac{d-2}{d-1}\widehat{R} =Rˇ^+\RomanbarII^μab\RomanbarII^μab.\displaystyle=\widehat{\check{R}}+\widehat{\Romanbar{II}}_{\mu ab}\widehat{\Romanbar{II}}^{\mu ab}. (276)

For codimension q=2q=2 the normal contraction of the Weyl tensor is equal to

Cjk=jk2Cvuvu,\displaystyle C_{jk}{}^{jk}=2C_{vuvu}, (277)

where {v,u}\{v,u\} is any orthonormal basis in the normal bundle. Therefore, for q=2q=2 we find

d3d1R^\displaystyle\frac{d-3}{d-1}\widehat{R} =Rˇ^+\RomanbarII^μab\RomanbarII^μab+2Cvuvu.\displaystyle=\widehat{\check{R}}+\widehat{\Romanbar{II}}_{\mu ab}\widehat{\Romanbar{II}}^{\mu ab}+2C_{vuvu}. (278)

Appendix B Conformal circle equation

In this appendix we rewrite the conformal circle in a form needed in section 2.3. The conformal circle equation is

taμ=3uau2aμ3a22u2uμ+u2uνPνμ2PαβuαuβUμ.\displaystyle\nabla_{t}a^{\mu}=\frac{3u\cdot a}{u^{2}}a^{\mu}-\frac{3a^{2}}{2u^{2}}u^{\mu}+u^{2}u^{\nu}P^{\mu}_{\nu}-2P_{\alpha\beta}u^{\alpha}u^{\beta}U^{\mu}. (279)

Define

f\displaystyle f =uau2,h=a2u2.\displaystyle=\frac{u\cdot a}{u^{2}},\quad h=\frac{a^{2}}{u^{2}}. (280)

We calculate

tf\displaystyle\nabla_{t}f =h2f2+utau2=12h+f2uμuνPμν,\displaystyle=h-2f^{2}+\frac{u\cdot\nabla_{t}a}{u^{2}}=-\tfrac{1}{2}h+f^{2}-u^{\mu}u^{\nu}P_{\mu\nu}, (281)
th\displaystyle\nabla_{t}h =2fh+2atau2=fh+2aμuνPμν4Pαβuαuβf.\displaystyle=-2fh+2\frac{a\cdot\nabla_{t}a}{u^{2}}=fh+2a^{\mu}u^{\nu}P_{\mu\nu}-4P_{\alpha\beta}u^{\alpha}u^{\beta}f. (282)

Altogether, we can write this as a system

tγμ\displaystyle\nabla_{t}\gamma^{\mu} =uμ,\displaystyle=u^{\mu}, (283)
tuμ\displaystyle\nabla_{t}u^{\mu} =aμ,\displaystyle=a^{\mu}, (284)
taμ\displaystyle\nabla_{t}a^{\mu} =3faμ32huμ+u2uνPνμ2Pαβuαuβuμ,\displaystyle=3fa^{\mu}-\tfrac{3}{2}hu^{\mu}+u^{2}u^{\nu}P^{\mu}_{\nu}-2P_{\alpha\beta}u^{\alpha}u^{\beta}u^{\mu}, (285)
tf\displaystyle\nabla_{t}f =12h+f2uμuνPμν,\displaystyle=-\tfrac{1}{2}h+f^{2}-u^{\mu}u^{\nu}P_{\mu\nu}, (286)
th\displaystyle\nabla_{t}h =fh+2aμuνPμν4Pαβuαuβf,\displaystyle=fh+2a^{\mu}u^{\nu}P_{\mu\nu}-4P_{\alpha\beta}u^{\alpha}u^{\beta}f, (287)

where the right-hand sides are now smooth functions of γ,u,a,f,h\gamma,u,a,f,h.

Appendix C Two-derivative identity operator couplings

In this section we classify two-derivative terms in the identity operator coupling. We first consider line defects, and then the codimension-1 and codimension-2 defects which are not line defects.

C.1 Line defects

Line defects have no intrinsic curvature invariants, and (82) implies that no extrinsic curvature invariants are available either (since the second fundamental form only has the trace component). Therefore, in building the effective action we can only use bulk curvature invariants and the derivatives of vμv^{\mu}. For bulk curvatures, it is convenient to decompose the Riemann tensor into Weyl tensor and Schouten tensor (266): Weyl tensor is Weyl-invariant, while Schouten tensor is constrained by (84). Note that the trace of P^\widehat{P} is proportional to the Ricci scalar.

d=3d=3 bulk

Recall that in section 3.6 we defined uμ=ε^μvμρσtσu^{\mu}=\widehat{\varepsilon}^{\mu}{}_{\rho\sigma}v^{\mu}t^{\sigma} where tμt^{\mu} is the unit vector (for fusion metric) tangent to the line defect.

We have only two ways to form 2-derivative invariants. The first is to use bulk curvature tensors, of which we only have the Schouten tensor PP available (Weyl tensor vanishes in d=3d=3). Due to (84) the only non-zero components are

P^tt,P^tv,P^tu,\displaystyle\widehat{P}_{tt},\quad\widehat{P}_{tv},\quad\widehat{P}_{tu}, (288)

of which P^tt\widehat{P}_{tt} can be expressed in terms of the Ricci scalar R^\widehat{R} due to P^vv=P^uu=0\widehat{P}_{vv}=\widehat{P}_{uu}=0.

The second option is to take covariant derivatives of vv. Up to total derivatives, the only two-derivative term is

^tv^tv=(utv)2.\displaystyle\widehat{\nabla}_{t}^{\perp}v\cdot\widehat{\nabla}_{t}^{\perp}v=(u\cdot\nabla_{t}^{\perp}v)^{2}. (289)

Overall, the 2-derivative contribution to the identity operator part of the effective action for line defects in d=3d=3 is

Seff(2)=\displaystyle S^{(2)}_{\text{eff}}= 𝑑zγ^(a2,1^tv^tv+a2,2R^+ia2,3P^tv+ia2,4P^tu).\displaystyle\int dz\sqrt{\widehat{\gamma}}\left(a_{2,1}\widehat{\nabla}_{t}^{\perp}v\cdot\widehat{\nabla}_{t}^{\perp}v+a_{2,2}\widehat{R}+ia_{2,3}\widehat{P}_{tv}+ia_{2,4}\widehat{P}_{tu}\right). (290)

The Schouten tensor in the above formula can be replaced with the Ricci tensor at the cost of rescaling the Wilson coefficients.

d4d\geq 4 bulk

Compared to the d=3d=3 case the only difference comes from the treatment of bulk curvature invariants. Firstly, for Schouten tensor, the only available components are P^tt\widehat{P}_{tt} and P^vt\widehat{P}_{vt}, but P^tt\widehat{P}_{tt} can again be expressed in terms of R^\widehat{R}.

Secondly, Weyl tensor is now non-zero. If d>4d>4 then the only way we can set indices is CvtvtC_{vtvt} due to the anti-symmetry properties. In d=4d=4 we can also form C~vtvt\widetilde{C}_{vtvt} where

C~μνσρ=12ϵ^μνCλκσρλκ.\displaystyle\widetilde{C}_{\mu\nu\sigma\rho}=\tfrac{1}{2}\widehat{\epsilon}_{\mu\nu}{}^{\lambda\kappa}C_{\lambda\kappa\sigma\rho}. (291)

Overall, the 2-derivative contribution to the identity operator part of the effective action for line defects in d4d\geq 4 is

Seff(2)=\displaystyle S^{(2)}_{\text{eff}}= 𝑑zγ^(a2,1^tv^tv+a2,2R^+ia2,3P^tv+a2,4Cvtvt+ia2,5C~vtvt),\displaystyle\int dz\sqrt{\widehat{\gamma}}\left(a_{2,1}\widehat{\nabla}_{t}^{\perp}v\cdot\widehat{\nabla}_{t}^{\perp}v+a_{2,2}\widehat{R}+ia_{2,3}\widehat{P}_{tv}+a_{2,4}C_{vtvt}+ia_{2,5}\widetilde{C}_{vtvt}\right), (292)

where the C~vtvt\widetilde{C}_{vtvt} has to be omitted for d>4d>4. Note that the terms containing the Weyl tensor are zero when the physical metric is conformally-flat. Again, the Schouten tensor in the above formula can be replaced with the Ricci tensor at the cost of rescaling the corresponding Wilson coefficient.

C.2 Codimension q=1q=1 defects (p>1p>1)

In the case of q=1q=1 the vector field vv is simply the unit normal to 𝒟1\mathcal{D}_{1} and thus its covariant derivatives vanish. If the defects 𝒟1\mathcal{D}_{1} or 𝒟2\mathcal{D}_{2} carry a normal bundle orientation, it might be possible to define a normal unit vector field nn on 𝒟1\mathcal{D}_{1} using these orientations. In this case, we can form nv=±1n\cdot v=\pm 1 and all Wilson coefficients may depend on this sign.

The independent components of the Schouten tensor that we can use at 2-derivative order are (taking into account (84))

P^va,P^ab,\displaystyle\widehat{P}_{va},\quad\widehat{P}_{ab}, (293)

where as always P^va=vμPμa\widehat{P}_{va}=v^{\mu}P_{\mu a} and a,ba,b are indices along the defect. However, we do not have any 0-derivative object to contract the open indices with. For similar reasons, in d4d\geq 4 we cannot form any terms from the Weyl tensor CC or (for d=4d=4) its dual C~\widetilde{C} defined in (291), or by using the first order covariant derivatives of \RomanbarII^abμ\widehat{\Romanbar{II}}^{\mu}_{ab}.

The two-derivative terms therefore have to be constructed from the bulk and defect Ricci scalars R^\widehat{R} and Rˇ^\widehat{\check{R}} or from two copies of \RomanbarII^abμ\widehat{\Romanbar{II}}^{\mu}_{ab}. Keeping in mind (276), we then find the following 2-derivative contribution to the identity operator part of the effective action for codimension q=1q=1 defects with p>1p>1,

Seff(2)=dpzγ^(a2,1R^+a2,2Rˇ^).\displaystyle S^{(2)}_{\text{eff}}=\int d^{p}z\sqrt{\widehat{\gamma}}\left(a_{2,1}\widehat{R}+a_{2,2}\widehat{\check{R}}\right). (294)

Note that when p=2p=2, the intrinsic Ricci scalar Rˇ^\widehat{\check{R}} is (locally) a total derivative and the corresponding term captures only the topological data about the defect.

C.3 Codimension q=2q=2 defects (p>1p>1)

When q=2q=2 we can define a normal vector uμ=ϵ^μvννu^{\mu}=\widehat{\epsilon}^{\mu}{}_{\nu}v^{\nu} using the normal bundle Levi-Civita tensor ϵ^μν\widehat{\epsilon}^{\mu}{}_{\nu}. The vectors vv and uu now span the normal bundle. Compared to q=1q=1 this allows us to define more terms using the second fundamental form,

(v\RomanbarII^ab)(v\RomanbarII^ab),(u\RomanbarII^ab)(u\RomanbarII^ab),(u\RomanbarII^ab)(v\RomanbarII^ab),(u\RomanbarII^ab)(v\RomanbarII^a)cϵ^bc,\displaystyle(v\cdot\widehat{\Romanbar{II}}_{ab})(v\cdot\widehat{\Romanbar{II}}^{ab}),\quad(u\cdot\widehat{\Romanbar{II}}_{ab})(u\cdot\widehat{\Romanbar{II}}^{ab}),\quad(u\cdot\widehat{\Romanbar{II}}_{ab})(v\cdot\widehat{\Romanbar{II}}^{ab}),\quad(u\cdot\widehat{\Romanbar{II}}_{ab})(v\cdot\widehat{\Romanbar{II}}^{a}{}_{c})\widehat{\epsilon}^{bc}, (295)

where in the last term ϵ^bc\widehat{\epsilon}^{bc} is the defect Levi-Civita tensor which is available only if d=4d=4 and p=2p=2. Equation (278) allows us to eliminate (u\RomanbarII^ab)(u\RomanbarII^ab)(u\cdot\widehat{\Romanbar{II}}_{ab})(u\cdot\widehat{\Romanbar{II}}^{ab}) using

(v\RomanbarII^ab)(u\RomanbarII^ab)+(v\RomanbarII^ab)(u\RomanbarII^ab)=\RomanbarII^μab\RomanbarII^μab.\displaystyle(v\cdot\widehat{\Romanbar{II}}_{ab})(u\cdot\widehat{\Romanbar{II}}^{ab})+(v\cdot\widehat{\Romanbar{II}}_{ab})(u\cdot\widehat{\Romanbar{II}}^{ab})=\widehat{\Romanbar{II}}_{\mu ab}\widehat{\Romanbar{II}}^{\mu ab}. (296)

Furthermore, terms involving Weyl tensor are now available. In d=4d=4 we can find two independent components (any other suitable components can be expressed in terms of these two)

Cvuvu,C~vuvu,\displaystyle C_{vuvu},\quad\widetilde{C}_{vuvu}, (297)

where C~\widetilde{C} is defined in (291). For d>4d>4 only CvuvuC_{vuvu} is available.

Finally, derivatives of vv are now nonzero. Since g^μνvμvν=1\widehat{g}_{\mu\nu}v^{\mu}v^{\nu}=1, the only non-trivial components are u^avu\cdot\widehat{\nabla}_{a}^{\perp}v and the only possible term (up to total derivatives) is

(u^av)(u^av)=(^av^av).\displaystyle(u\cdot\widehat{\nabla}_{a}^{\perp}v)(u\cdot\widehat{\nabla}^{\perp a}v)=(\widehat{\nabla}_{a}^{\perp}v\cdot\widehat{\nabla}^{\perp a}v). (298)

Overall we find the following 2-derivative contribution to the identity operator part of the effective action for codimension q=2q=2 defects with p>1p>1,

Seff(2)=dzγ^(\displaystyle S^{(2)}_{\text{eff}}=\int dz\sqrt{\widehat{\gamma}}\Big{(} a2,1(v\RomanbarII^ab)(v\RomanbarII^ab)+a2,2(u\RomanbarII^ab)(v\RomanbarII^ab)+ia2,3(u\RomanbarII^ab)(v\RomanbarII^a)cϵ^bc\displaystyle a_{2,1}(v\cdot\widehat{\Romanbar{II}}_{ab})(v\cdot\widehat{\Romanbar{II}}^{ab})+a_{2,2}(u\cdot\widehat{\Romanbar{II}}_{ab})(v\cdot\widehat{\Romanbar{II}}^{ab})+ia_{2,3}(u\cdot\widehat{\Romanbar{II}}_{ab})(v\cdot\widehat{\Romanbar{II}}^{a}{}_{c})\widehat{\epsilon}^{bc}
+a2,4(^av^av)+a2,5R^+a2,6Rˇ^+a2,7Cvuvu+ia2,8C~vuvu).\displaystyle+a_{2,4}(\widehat{\nabla}_{a}^{\perp}v\cdot\widehat{\nabla}^{\perp a}v)+a_{2,5}\widehat{R}+a_{2,6}\widehat{\check{R}}+a_{2,7}C_{vuvu}+ia_{2,8}\widetilde{C}_{vuvu}\Big{)}. (299)

The terms involving C~vuvu\widetilde{C}_{vuvu} and (u\RomanbarII^ab)(v\RomanbarII^a)cϵ^bc(u\cdot\widehat{\Romanbar{II}}_{ab})(v\cdot\widehat{\Romanbar{II}}^{a}{}_{c})\widehat{\epsilon}^{bc} are to be omitted for d>4d>4. Note that when p=2p=2, the intrinsic Ricci scalar Rˇ^\widehat{\check{R}} is (locally) a total derivative and the corresponding term captures only the topological data about the defect.

Appendix D Ansatz for the Weyl anomaly of a surface defect in d=4d=4

The initial ansatz for the defect contribution to the Weyl anomaly of a surface defect in a d=4d=4 bulk, as used in section 6.4 is given by:

𝒜𝒟(1)(g,ω)=\displaystyle\mathcal{A}^{(1)}_{\mathcal{D}}(g,\omega)= d2zγi=122fiFi(ω),\displaystyle\int d^{2}z\sqrt{\gamma}\sum_{i=1}^{22}f_{i}F_{i}(\omega), (300)

where:

  • F1(ω)=\RomanbarIIχ2ωF_{1}(\omega)=\Romanbar{II}_{\chi}^{2}\omega,

  • F2(ω)=\RomanbarIIχ¯2ωF_{2}(\omega)=\Romanbar{II}_{\overline{\chi}}^{2}\omega,

  • F3(ω)=\RomanbarIIχ\RomanbarIIχ¯ωF_{3}(\omega)=\Romanbar{II}_{\chi}\Romanbar{II}_{\overline{\chi}}\omega,

  • F4(ω)=\RomanbarII̊χζζ\RomanbarII̊χζ¯ζ¯ωF_{4}(\omega)=\mathring{\Romanbar{II}}_{\chi\zeta\zeta}\mathring{\Romanbar{II}}_{\chi\overline{\zeta}\overline{\zeta}}\omega,

  • F5(ω)=\RomanbarII̊χζζ\RomanbarII̊χ¯ζ¯ζ¯ωF_{5}(\omega)=\mathring{\Romanbar{II}}_{\chi\zeta\zeta}\mathring{\Romanbar{II}}_{\overline{\chi}\overline{\zeta}\overline{\zeta}}\omega,

  • F6(ω)=\RomanbarII̊χ¯ζζ\RomanbarII̊χζ¯ζ¯ωF_{6}(\omega)=\mathring{\Romanbar{II}}_{\overline{\chi}\zeta\zeta}\mathring{\Romanbar{II}}_{\chi\overline{\zeta}\overline{\zeta}}\omega,

  • F7(ω)=\RomanbarII̊χ¯ζζ\RomanbarII̊χ¯ζ¯ζ¯ωF_{7}(\omega)=\mathring{\Romanbar{II}}_{\overline{\chi}\zeta\zeta}\mathring{\Romanbar{II}}_{\overline{\chi}\overline{\zeta}\overline{\zeta}}\omega,

  • F8(ω)=RˇωF_{8}(\omega)=\check{R}\omega,

  • F9(ω)=RχχωF_{9}(\omega)=R_{\chi\chi}\omega,

  • F10(ω)=RχχωF_{10}(\omega)=R_{\chi\chi}\omega,

  • F11(ω)=Rχ¯χ¯ωF_{11}(\omega)=R_{\overline{\chi}\overline{\chi}}\omega,

  • F12(ω)=Cχχ¯χχ¯ωF_{12}(\omega)=C_{\chi\overline{\chi}\chi\overline{\chi}}\omega,

  • F13(ω)=Cχχ¯ζζ¯ωF_{13}(\omega)=C_{\chi\overline{\chi}\zeta\overline{\zeta}}\omega,

  • F14(ω)=χμg¯ababχμωF_{14}(\omega)=\chi_{\mu}\overline{g}^{ab}\nabla_{a}\nabla^{\perp}_{b}\chi^{\mu}\omega,

  • F15(ω)=χ¯μg¯ababχμωF_{15}(\omega)=\overline{\chi}_{\mu}\overline{g}^{ab}\nabla_{a}\nabla^{\perp}_{b}\chi^{\mu}\omega,

  • F16(ω)=\RomanbarIIχχωF_{16}(\omega)=\Romanbar{II}_{\chi}\partial_{\chi}\omega,

  • F17(ω)=\RomanbarIIχχ¯ωF_{17}(\omega)=\Romanbar{II}_{\chi}\partial_{\overline{\chi}}\omega,

  • F18(ω)=\RomanbarIIχ¯χωF_{18}(\omega)=\Romanbar{II}_{\overline{\chi}}\partial_{\chi}\omega,

  • F19(ω)=\RomanbarIIχ¯χ¯ωF_{19}(\omega)=\Romanbar{II}_{\overline{\chi}}\partial_{\overline{\chi}}\omega,

  • F20(ω)=χμχνμνωF_{20}(\omega)=\chi^{\mu}\chi^{\nu}\nabla_{\mu}\partial_{\nu}\omega,

  • F21(ω)=χμχ¯νμνωF_{21}(\omega)=\chi^{\mu}\overline{\chi}^{\nu}\nabla_{\mu}\partial_{\nu}\omega,

  • F22(ω)=χ¯μχ¯νμνωF_{22}(\omega)=\overline{\chi}^{\mu}\overline{\chi}^{\nu}\nabla_{\mu}\partial_{\nu}\omega.

References