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

Massless limit and conformal soft limit for celestial massive amplitudes

Wei Fan Department of Physics, School of Science, Jiangsu University of Science and Technology, Zhenjiang, 212100, China [email protected]
Abstract

In celestial holography, the massive and massless scalars in 4d space-time are represented by the Fourier transform of the bulk-to-boundary propagators and the Mellin transform of plane waves respectively. Recently, the 3pt celestial amplitude of one massive scalar and two massless scalars was discussed in arXiv:2312.08597. In this paper, we compute the 3pt celestial amplitude of two massive scalars and one massless scalar. Then we take the massless limit m0m\to 0 for one of the massive scalars, during which process the gamma function Γ(1Δ)\Gamma(1-\Delta) appears. By requiring the resulting amplitude to be well-defined, that is it goes to the 3pt amplitude of arXiv:2312.08597, the scaling dimension of this massive scalar has to be conformally soft Δ1\Delta\to 1 (note that in this paper the ’conformally soft’ is associated with the massless scalar after taking the massless limit. For the massive scalar, Δ1\Delta\to 1 is only choosing a specific dimension, but after the massless limit m0m\to 0 it becomes the conformal soft dimension. Because the specific choice of dimension and the massless limit are taken together in this paper, we will directly use ’conformally soft’ for simplicity.). The pole 1/(1Δ)1/(1-\Delta) coming from Γ(1Δ)\Gamma(1-\Delta) is crucial for this massless limit. Without it the resulting amplitude would be zero. This can be compared with the conformal soft limit in celestial gluon amplitudes, where a singularity 1/(Δ1)1/(\Delta-1) arises and the leading contribution comes from the soft energy ω0\omega\to 0. The phase factors in the massless limit of massive conformal primary wave functions, dicussed in arXiv:1705.01027, plays an import and consistent role in the celestial massive amplitudes. Furthermore, the subleading orders m2nm^{2n} can also contribute poles when the scaling dimension is analytically continued to Δ=1n\Delta=1-n or Δ=2\Delta=2, and we find that this consistent massless limit only exists for dimensions belonging to the generalized conformal primary operators Δ20\Delta\in 2-\mathbb{Z}_{\geqslant 0} of massless bosons.

I Introduction

Celestial conformal field theory (CCFT) is a flat holography [1, 2, 3, 4] of Minkowski spacetime. The bulk QFT in Minkowski spacetime 1,3\mathbb{R}^{1,3} can be recast as a boundary CFT in the celestial sphere. The amplitudes are converted from the momentum basis to the boost basis (the so-called conformal primary wavefunction) [5] and then become the CFT correlation functions. The consequence of 4d soft theorem to celestial holography is one of the import problems. For example, the soft energy ω0\omega\to 0 generates the leading contribution to the conformal soft limit Δ1\Delta\to 1 in celestial gluon amplitudes [6, 7, 8, 9].

A natural question to ask is what would be the corresponding physics related with the conformal soft limit of celestial massive amplitudes? We address this question in this paper using the 3pt celestial amplitude of two massive scalars and one massless scalar. We find that the massless limit m0m\to 0 of this amplitude leads to the conformal soft limit Δ1\Delta\to 1, to make the resulting amplitude well-defined. On the basis level, the massless limit of the conformal primary wave functions has a phase factor mΔm^{-\Delta} [5]. After going to the celestial amplitudes, the massless limit is accompanied by the gamma function Γ(1Δ)\Gamma(1-\Delta), which contributes a pole in the conformal soft limit Δ1\Delta\to 1. This resembles, in spirit, to the conformal soft limit of celestial gluons, which appears in their amplitudes rather than conformal primary waves.

The celestial massive amplitudes are hard to compute in 4d space-time. In the original paper of massive conformal primary waves [10], the 3pt celestial amplitudes of three massive scalars was computed in the near-extremal limit m1=2(1+ϵ)m,m2=m3=mm_{1}=2(1+\epsilon)m,m_{2}=m_{3}=m. The result is the tree-level 3pt Witten diagram in the leading order of ϵ\sqrt{\epsilon} when ϵ0\epsilon\to 0. The 3pt celestial amplitude of one massive scalar and two massless scalars ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle was computed recently by [11]111This kind of celestial amplitudes was firstly discussed by  [12] in 3d space-time, for the implications of the 4d optical theorem on celestial holography., where the non-local behavior of massive states on celestial sphere was discussed. Since massive states are highly nontrivial in celestial holography, it is meaningful to investigate the physics related with their conformal soft limit. Here we continue to compute the 3pt celestial amplitude of two massive scalars and one massless scalar ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle. Using the standard method of hyperbolic coordinates and Feynman parametrization, the two masses m1m_{1} and m2m_{2} appear only in the finite integral of the Feynman parameters. Then we can safely take the massless limit m20m_{2}\to 0. Combined with the conformal soft limit Δ21\Delta_{2}\to 1, the 3pt massive amplitude ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle has the desired limit ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle.

On the basis level, the massless limit of the massive conformal primary wave function leads to two massless fields ΦΔ2mϕΔ2,\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS\Phi_{\Delta_{2}}^{m}\to\phi_{\Delta_{2}},\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}} [5], which are the same for Δ2=1\Delta_{2}=1. They are different for general values of Δ2\Delta_{2}. By studying subleading orders m22nm_{2}^{2n} of the massless limit, we find that it leads to the 3pt amplitude ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle for the analytically continued dimension Δ2=1n\Delta_{2}=1-n, and to the 3pt amplitude ΦΔ1m1\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTSϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}\phi_{\Delta_{3}}\rangle for the analytically continued dimension Δ2=2\Delta_{2}=2. For analytically continued dimension Δ23\Delta_{2}\geq 3, there is a massless limit but it is neither ΦΔ1m1\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTSϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}\phi_{\Delta_{3}}\rangle nor ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle. So it is only for dimensions Δ22\Delta_{2}\leq 2 that the massless limit of this amplitude is consistent. These dimensions happen to belong to the range of generalized conformal primary operators Δ20\Delta\in 2-\mathbb{Z}_{\geqslant 0} of massless bosons [13]. So the consistent massless limit of ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle picks up these generalized conformal primary operators. Note that this amplitude was also discussed recently in [14], whose results and physics content have no overlap with us.

The paper is organized as follows. In Section 2, we give a short account of the notation and briefly review the 3pt celestial amplitude ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle. Then we compute the shadowed 3pt amplitude ΦΔ1m1\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTSϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}\phi_{\Delta_{3}}\rangle. In Section 3, we compute the 3pt celestial amplitude ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle using the Feynman parametrization. The complete result of this amplitude is put to the appendix. In Section 4, we investigate the massless limit m20m_{2}\to 0 and the conformal soft limit Δ21\Delta_{2}\to 1. Then we discuss the subleading terms m22nm_{2}^{2n} and the analytic continuation to Δ2=1n\Delta_{2}=1-n and Δ2=2\Delta_{2}=2. We conclude with a discussion of open questions in Section 5.

II Preliminaries

In CCFT, the massless four-momentum piμ=ϵiωiq^iμp_{i}^{\mu}=\epsilon_{i}\omega_{i}\hat{q}_{i}^{\mu} is parameterized by the energy ωi\omega_{i} and the point (wi,w¯i)(w_{i},\bar{w}_{i}) of the celestial sphere via the formula [5]

q^i=(1+wiw¯i,wi+w¯i,i(wiw¯i),1wiw¯i),\displaystyle\hat{q}_{i}=(1+w_{i}\bar{w}_{i},w_{i}+\bar{w}_{i},-i(w_{i}-\bar{w}_{i}),1-w_{i}\bar{w}_{i}), (II.1)

where ϵi=±1\epsilon_{i}=\pm 1 represents outgoing/incoming momentum. During the computation, we use the coordinate wi=(Re(wi),Im(wi))\vec{w}_{i}=(\operatorname{Re}(w_{i}),\operatorname{Im}(w_{i})) for the celestial sphere, in order to save space. Each 4d massless scalar corresponds to a scalar conformal primary wave function ϕΔ(Xμ;w)\phi_{\Delta}(X^{\mu};\vec{w}) (the so-called boost basis) via the Mellin transform

ϕΔ(±)(Xμ;w)0𝑑ωωΔ1e±iωq^(w)Xεω,\displaystyle\phi_{\Delta}^{(\pm)}(X^{\mu};\vec{w})\coloneqq\int_{0}^{\infty}d\omega\omega^{\Delta-1}e^{\pm i\omega\hat{q}(\vec{w})\cdot X-\varepsilon\omega}, (II.2)

where ε>0\varepsilon>0 is a regularization parameter and the conformal dimensions are hi=h¯i=Δi/2=(1+iλi)/2,λih_{i}=\bar{h}_{i}=\Delta_{i}/2=(1+i\lambda_{i})/2,\lambda_{i}\in\mathbb{R}.

The massive four-momentum piμ=ϵimip^μp_{i}^{\mu}=\epsilon_{i}m_{i}\hat{p}^{\mu} is parameterized by the mass of the particle mim_{i} and the hyperbolic coordinate (yi,zi)(y_{i},\vec{z}_{i}) via the formula [10]

p^iμ(yi,zi)=(1+yi2+|zi|22yi,ziyi,1yi2|zi|22yi).\displaystyle\hat{p}_{i}^{\mu}(y_{i},\vec{z}_{i})=\left(\frac{1+y_{i}^{2}+|z_{i}|^{2}}{2y_{i}},\frac{\vec{z}_{i}}{y_{i}},\frac{1-y_{i}^{2}-|z_{i}|^{2}}{2y_{i}}\right). (II.3)

The massive scalar corresponds to a scalar conformal primary wave ΦΔm(Xμ;w)\Phi_{\Delta}^{m}(X^{\mu};\vec{w}) via the Fourier transform

ΦΔm(±)(Xμ;w)0dyy3d2zGΔ(y,z;w)e±imp^(y,z)X,\displaystyle\Phi_{\Delta}^{m(\pm)}(X^{\mu};\vec{w})\coloneqq\int_{0}^{\infty}\frac{dy}{y^{3}}\int d^{2}zG_{\Delta}(y,\vec{z};\vec{w})e^{\pm im\hat{p}(y,\vec{z})\cdot X}, (II.4)

where GΔ(y,z;w)G_{\Delta}(y,\vec{z};\vec{w}) is the bulk-to-boundary propagator with scaling dimension Δ\Delta

GΔ(p^(y,z);q^(w))=1(p^q^)Δ=(yy2+|zw|2)Δ.\displaystyle G_{\Delta}(\hat{p}(y,\vec{z});\hat{q}(\vec{w}))=\frac{1}{(-\hat{p}\cdot\hat{q})^{\Delta}}=\left(\frac{y}{y^{2}+|\vec{z}-\vec{w}|^{2}}\right)^{\Delta}. (II.5)

The Lorentz transformation ΛνμSO(1,3)\Lambda^{\mu}_{\nu}\in SO(1,3) acts non-linearly on the coordinates ww(w),zz(y,z),yy(y,z)\vec{w}\to\vec{w}^{\prime}(\vec{w}),\vec{z}\to\vec{z}^{\prime}(y,\vec{z}),y\to y^{\prime}(y,\vec{z}), but the four-momenta transform linearly

p^μ(y,z)=Λμp^νν,qμ(w)=|ww|1/2Λνμqν(w).\displaystyle\hat{p}^{\mu}\left(y^{\prime},\vec{z}^{\prime}\right)=\Lambda^{\mu}{}_{\nu}\hat{p}^{\nu},\,q^{\mu}\left(\vec{w}^{\prime}\right)=\left|\frac{\partial\vec{w}^{\prime}}{\partial\vec{w}}\right|^{1/2}\Lambda_{\nu}^{\mu}q^{\nu}(\vec{w}). (II.6)

Then the conformal primary waves have the conformal symmetry on the celestial sphere

{ΦΔm,ϕΔ}(ΛνμXν;w(w))=|ww|Δ/2{ΦΔm,ϕΔ}(Xμ;w),\displaystyle\Big{\{}\Phi_{\Delta}^{m},\phi_{\Delta}\Big{\}}(\Lambda_{\nu}^{\mu}X^{\nu};\vec{w}^{\prime}(\vec{w}))=\left|\frac{\partial\vec{w}^{\prime}}{\partial\vec{w}}\right|^{-\Delta/2}\Big{\{}\Phi_{\Delta}^{m},\phi_{\Delta}\Big{\}}(X^{\mu};\vec{w}), (II.7)

where we write ΦΔm(X;w)\Phi_{\Delta}^{m}(X;\vec{w}) and ϕΔ(X;w)\phi_{\Delta}(X;\vec{w}) together to save space.

In the massless limit, the massive conformal primary wavefunction ΦΔm(±)(Xμ;w)\Phi_{\Delta}^{m(\pm)}(X^{\mu};\vec{w}) has the following behavior [5]

ΦΔm(X;w)m0(m2)ΔπΓ(Δ1)Γ(Δ)ϕΔ(Xμ;w)+(m2)Δ~πΓ(Δ~1)Γ(Δ~)\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~\stretchto\scaleto\SavedStyle.54670.5OcFTS(Xμ;w)\displaystyle\Phi_{\Delta}^{m}(X;\vec{w})\overset{m\rightarrow 0}{\longrightarrow}\left(\frac{m}{2}\right)^{-\Delta}\frac{\pi\Gamma(\Delta-1)}{\Gamma\left(\Delta\right)}\phi_{\Delta}(X^{\mu};\vec{w})+\left(\frac{m}{2}\right)^{-\tilde{\Delta}}\frac{\pi\Gamma(\tilde{\Delta}-1)}{\Gamma\left(\tilde{\Delta}\right)}\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}(X^{\mu};\vec{w})

where Δ~=2Δ\tilde{\Delta}=2-\Delta is the shadow dimension and \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}} is the shadow operator of ϕΔ~\phi_{\tilde{\Delta}}. In the basis level, the massless limit is not well-defined, due to this phase factor mΔm^{-\Delta}. In the amplitude level, however, things are different. The physical interaction between particles make the massless limit well-defined, provided that the conformal soft limit Δ1\Delta\to 1 is together with the massless limit m0m\to 0. This is what we will show in this paper. It turns out that the phase factor mΔm^{-\Delta} is crucial for the massless limit of celestial amplitudes.

II.1 The 3pt celestial amplitude ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle and ΦΔ1m1ϕΔ~2~ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\widetilde{\phi_{\tilde{\Delta}_{2}}}\phi_{\Delta_{3}}\rangle

Here we briefly review the 3pt celestial amplitude ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle appearing in [11]. This 3pt amplitude is given by the following integral

ΦΔ1m1(w1)ϕΔ2(w2)ϕΔ3(w3)=iλd4XΦΔ1m1()(Xμ;w1)i=23ϕΔi(+)(Xμ;wi),\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}(\vec{w}_{1})\phi_{\Delta_{2}}(\vec{w}_{2})\phi_{\Delta_{3}}(\vec{w}_{3})\rangle=i\lambda\int d^{4}X\,\Phi_{\Delta_{1}}^{m_{1}(-)}\left(X^{\mu};\vec{w}_{1}\right)\prod_{i=2}^{3}\phi_{\Delta_{i}}^{(+)}\left(X^{\mu};\vec{w}_{i}\right), (II.8)

where ΦΔimi(wi)\Phi_{\Delta_{i}}^{m_{i}}(\vec{w}_{i}) are the corresponding conformal primary operators in CCFT. The integration over XX can be computed first, which leads to a momentum conservating delta function

dy1y13d2z1(y1y12+|z1w1|2)Δ1(i=23dωiωiωiΔi)δ4(m1p^1+ω2q^2+ω3q^3).\displaystyle\int\frac{dy_{1}}{y_{1}^{3}}\int d^{2}z_{1}\left(\frac{y_{1}}{y_{1}^{2}+|\vec{z}_{1}-\vec{w}_{1}|^{2}}\right)^{\Delta_{1}}\int\left(\prod_{i=2}^{3}\frac{d\omega_{i}}{\omega_{i}}\omega_{i}^{\Delta_{i}}\right)\delta^{4}(-m_{1}\hat{p}_{1}+\omega_{2}\hat{q}_{2}+\omega_{3}\hat{q}_{3}). (II.9)

This momentum conservating delta function can be used to fix the integration variables (y1,z1,ω2)(y_{1},\vec{z}_{1},\omega_{2}), then only one integration ω3\omega_{3} is left, which can be done analytically. The final result is

iλ(2π)416(m12)Δ2+Δ34Γ(Δ1+Δ2Δ32)Γ(Δ1Δ2+Δ32)/Γ(Δ1)|w12|Δ1+Δ2Δ3|w13|Δ1+Δ3Δ2|w23|Δ2+Δ3Δ1,\displaystyle\frac{i\lambda(2\pi)^{4}}{16}\left(\frac{m_{1}}{2}\right)^{\Delta_{2}+\Delta_{3}-4}\frac{\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2})\Gamma(\frac{\Delta_{1}-\Delta_{2}+\Delta_{3}}{2})/\Gamma(\Delta_{1})}{|\vec{w}_{12}|^{\Delta_{1}+\Delta_{2}-\Delta_{3}}|\vec{w}_{13}|^{\Delta_{1}+\Delta_{3}-\Delta_{2}}|\vec{w}_{23}|^{\Delta_{2}+\Delta_{3}-\Delta_{1}}}, (II.10)

where wijwiwj\vec{w}_{ij}\coloneqq\vec{w}_{i}-\vec{w}_{j} and the overall constant are adapted to the notation of this paper.

In the following, we continue to compute the 3pt amplitude with two massive scalars and one massless scalar ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle. By taking one massive scalar to be massless m20m_{2}\to 0, we show that it reduces to the 3pt amplitude ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle as expected. The massless limit of ΦΔ2m(X;w)\Phi_{\Delta_{2}}^{m}(X;\vec{w}) contains both ϕΔ2\phi_{\Delta_{2}} and \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}, so we also need the 3pt amplitude ΦΔ1m1\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTSϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}\phi_{\Delta_{3}}\rangle. To do that, we need the shadow transformation.

The shadow of a primary operator ϕΔ(z,z¯)\phi_{\Delta}(z,\bar{z}) with conformal dimension h,h¯h,\bar{h}, hence with the scaling dimension Δ=h+h¯\Delta=h+\bar{h} and spin J=hh¯J=h-\bar{h}, is defined [15] as

ϕ~Δ~(z,z¯)\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ(y,y¯)\stretchto\scaleto\SavedStyle.54670.5OcFTS=kh,h¯d2y(zy)2h2(z¯y¯)2h¯2ϕΔ(y,y¯)\displaystyle\tilde{\phi}_{\tilde{\Delta}}(z,\bar{z})\coloneqq\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\Delta}(y,\bar{y})$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}=k_{h,\bar{h}}\int d^{2}y(z-y)^{2h-2}(\bar{z}-\bar{y})^{2\bar{h}-2}\phi_{\Delta}(y,\bar{y})

where the constant kh,h¯=(1)2(hh¯)Γ(22h)/(πΓ(2h¯1))k_{h,\bar{h}}=(-1)^{2(h-\bar{h})}\Gamma(2-2h)/(\pi\Gamma(2\bar{h}-1)) is chosen in such a way that, for integer or half-integer spin, ϕ~~(z,z¯)=ϕ(z,z¯)\tilde{\tilde{\phi}}(z,\bar{z})=\phi(z,\bar{z}). The shadow field ϕ~Δ~(z,z¯)\tilde{\phi}_{\tilde{\Delta}}(z,\bar{z}) is a primary operator with conformal dimension 1h,1h¯1-h,1-\bar{h}, hence with Δ~=2Δ\tilde{\Delta}=2-\Delta and J~=J\tilde{J}=-J.

Now we can perform the shadow transform on ϕΔ~2\phi_{\tilde{\Delta}_{2}} with conformal dimension h=h¯=Δ~2/2h=\bar{h}=\tilde{\Delta}_{2}/2 to get the shadowed field \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS(w2)=\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2(w2)\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}(\vec{w}_{2})=\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}(\vec{w}^{\prime}_{2})$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}

ΦΔ1m1(w1)\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS(w2)ϕΔ3(w3)=kh,h¯d2w2|w2w2|42Δ~2ΦΔ1m1(w1)ϕΔ~2(w2)ϕΔ3(w3)\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}(\vec{w}_{1})\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}(\vec{w}_{2})\phi_{\Delta_{3}}(\vec{w}_{3})\rangle=k_{h,\bar{h}}\int\frac{d^{2}\vec{w}^{\prime}_{2}}{|\vec{w}^{\prime}_{2}-\vec{w}_{2}|^{4-2\tilde{\Delta}_{2}}}\langle\Phi_{\Delta_{1}}^{m_{1}}(\vec{w}_{1})\phi_{\tilde{\Delta}_{2}}(\vec{w}^{\prime}_{2})\phi_{\Delta_{3}}(\vec{w}_{3})\rangle

Using the conformal symmetry, we shall perform a conformal transformation to set w10,w21,w3\vec{w}_{1}\to 0,\vec{w}_{2}\to 1,\vec{w}_{3}\to\infty, w2x\vec{w}^{\prime}_{2}\to\vec{x} and obtain the function

ΦΔ1m1(w1)\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}(\vec{w}_{1}) \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS(w2)ϕΔ3(w3)=\displaystyle\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}(\vec{w}_{2})\phi_{\Delta_{3}}(\vec{w}_{3})\rangle=
×Γ(2Δ~2)Γ(Δ1+Δ~2Δ32)Γ(Δ1Δ~2+Δ32)πΓ(Δ~21)Γ(Δ1)d2x|x|Δ3Δ1Δ~2|x1|2Δ~24.\displaystyle\times\frac{\Gamma(2-\tilde{\Delta}_{2})\Gamma(\frac{\Delta_{1}+\tilde{\Delta}_{2}-\Delta_{3}}{2})\Gamma(\frac{\Delta_{1}-\tilde{\Delta}_{2}+\Delta_{3}}{2})}{\pi\Gamma(\tilde{\Delta}_{2}-1)\Gamma(\Delta_{1})}\int\,d^{2}\vec{x}\,|\vec{x}|^{\Delta_{3}-\Delta_{1}-\tilde{\Delta}_{2}}|\vec{x}-1|^{2\tilde{\Delta}_{2}-4}. (II.11)

Such kind of 2d scalar integrals can be evaluated by analytic continuation of xx on the complex plane

d2x|x|2a|x1|2b=πΓ(1+a)Γ(1+b)Γ(1ab)Γ(a)Γ(b)Γ(2+a+b),\displaystyle\int\,d^{2}\vec{x}\,|\vec{x}|^{2a}|\vec{x}-1|^{2b}=\pi\frac{\Gamma(1+a)\Gamma(1+b)\Gamma(-1-a-b)}{\Gamma(-a)\Gamma(-b)\Gamma(2+a+b)}, (II.12)

the details of which can be seen in Chapter 7 of [16].

So we finally obtain the 3pt amplitude

ΦΔ1m1(w1)\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS(w2)ϕΔ3(w3)=iλ(2π)4m1Δ~2+Δ342Δ~2+Δ3Γ(Δ1+Δ2Δ32)Γ(Δ2+Δ3Δ12)Γ(Δ1)Γ(1+Δ1+Δ2+Δ32)Γ(1+Δ3Δ1Δ22)|w12|Δ1+Δ2Δ3|w13|Δ1+Δ3Δ2|w23|Δ2+Δ3Δ1\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}(\vec{w}_{1})\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}(\vec{w}_{2})\phi_{\Delta_{3}}(\vec{w}_{3})\rangle=\frac{i\lambda(2\pi)^{4}m_{1}^{\tilde{\Delta}_{2}+\Delta_{3}-4}}{2^{\tilde{\Delta}_{2}+\Delta_{3}}}\frac{\frac{\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2})\Gamma(\frac{\Delta_{2}+\Delta_{3}-\Delta_{1}}{2})}{\Gamma(\Delta_{1})}\frac{\Gamma(-1+\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2})}{\Gamma(1+\frac{\Delta_{3}-\Delta_{1}-\Delta_{2}}{2})}}{|\vec{w}_{12}|^{\Delta_{1}+\Delta_{2}-\Delta_{3}}|\vec{w}_{13}|^{\Delta_{1}+\Delta_{3}-\Delta_{2}}|\vec{w}_{23}|^{\Delta_{2}+\Delta_{3}-\Delta_{1}}}

III The 3pt celestial amplitude ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle

For convenience we choose the 1st particle to be incoming, the 2nd and the 3rd to be outgoing. The 3pt celestial amplitude ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle is given by the following integral

ΦΔ1m1(w1)ΦΔ2m2(w2)ϕΔ3(w3)=iλd4XΦΔ1m1()(X;w1)ΦΔ2m2(+)(X;w2)ϕΔ3(+)(X;w3).\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}(\vec{w}_{1})\Phi_{\Delta_{2}}^{m_{2}}(\vec{w}_{2})\phi_{\Delta_{3}}(\vec{w}_{3})\rangle=i\lambda\int d^{4}X\,\Phi_{\Delta_{1}}^{m_{1}(-)}\left(X;\vec{w}_{1}\right)\Phi_{\Delta_{2}}^{m_{2}(+)}\left(X;\vec{w}_{2}\right)\phi_{\Delta_{3}}^{(+)}\left(X;\vec{w}_{3}\right). (III.1)

After the integration over XX it becomes

iλ(2π)4(i=12dyid2ziyi3(yiyi2+|ziwi|2)Δi)dω3ω3ω3Δ3δ4(m1p^1+m2p^2+ω3q^3).\displaystyle i\lambda(2\pi)^{4}\displaystyle{\int\left(\prod_{i=1}^{2}\frac{dy_{i}d^{2}z_{i}}{y_{i}^{3}}\left(\frac{y_{i}}{y_{i}^{2}+|\vec{z}_{i}-\vec{w}_{i}|^{2}}\right)^{\Delta_{i}}\right)\int\frac{d\omega_{3}}{\omega_{3}}\omega_{3}^{\Delta_{3}}\delta^{4}(-m_{1}\hat{p}_{1}+m_{2}\hat{p}_{2}+\omega_{3}\hat{q}_{3})}. (III.2)

The momentum conservating delta function can be evaluated as

δ4(m1p^1+m2p^2+ω3q^3)=y14(m12y22+m22(z2w3)2)m13m22(y22+(z2w3)2)2δ(y1y1)δ2(z1z1)δ(ω3ω3),\displaystyle\delta^{4}(-m_{1}\hat{p}_{1}+m_{2}\hat{p}_{2}+\omega_{3}\hat{q}_{3})=\frac{y_{1}^{4}(m_{1}^{2}y_{2}^{2}+m_{2}^{2}(\vec{z}_{2}-\vec{w}_{3})^{2})}{m_{1}^{3}m_{2}^{2}(y_{2}^{2}+(\vec{z}_{2}-\vec{w}_{3})^{2})^{2}}\delta(y_{1}-y_{1}^{*})\delta^{2}(\vec{z}_{1}-\vec{z}_{1}^{*})\delta(\omega_{3}-\omega_{3}^{*}), (III.3)

where the factor is the Jacobian coming from solving the momentum conservation using hyperbolic coordinates. The fixed integration variables are as following

y1\displaystyle y_{1}^{*} =m1m2y2(y22+(z2w3)2)m12y22+m22(z2w3)2\displaystyle=\frac{m_{1}m_{2}y_{2}(y_{2}^{2}+(\vec{z}_{2}-\vec{w}_{3})^{2})}{m_{1}^{2}y_{2}^{2}+m_{2}^{2}(\vec{z}_{2}-\vec{w}_{3})^{2}}
z1\displaystyle\vec{z}_{1}^{*} =m22(y22+(z2w3)2)z2+y22(m12m22)w3m12y22+m22(z2w3)2\displaystyle=\frac{m_{2}^{2}(y_{2}^{2}+(\vec{z}_{2}-\vec{w}_{3})^{2})\vec{z}_{2}+y_{2}^{2}(m_{1}^{2}-m_{2}^{2})\vec{w}_{3}}{m_{1}^{2}y_{2}^{2}+m_{2}^{2}(\vec{z}_{2}-\vec{w}_{3})^{2}}
ω3\displaystyle\omega_{3}^{*} =y2(m12m22)2m2(y22+(z2w3)2).\displaystyle=\frac{y_{2}(m_{1}^{2}-m_{2}^{2})}{2m_{2}(y_{2}^{2}+(\vec{z}_{2}-\vec{w}_{3})^{2})}. (III.4)

Since y0,ω0y\geq 0,\omega\geq 0, the kinematics constrains the masses to be m1m2m_{1}\geq m_{2}.

This integral can be simplified by the conformal symmetry (II.7). This conformal symmetry fixes the w\vec{w} dependence of the integral to be

ΦΔ1m1ΦΔ2m2ϕΔ31|w12|Δ1+Δ2Δ3|w13|Δ1+Δ3Δ2|w23|Δ2+Δ3Δ1.\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle\propto\frac{1}{|\vec{w}_{12}|^{\Delta_{1}+\Delta_{2}-\Delta_{3}}|\vec{w}_{13}|^{\Delta_{1}+\Delta_{3}-\Delta_{2}}|\vec{w}_{23}|^{\Delta_{2}+\Delta_{3}-\Delta_{1}}}. (III.5)

The exact factor of this proportion can be computed by using the famous linear fractional transformation to fix the three points on the celestial sphere

w1,w21=(1,0),w30.\displaystyle\vec{w}_{1}\to\infty,\,\vec{w}_{2}\to\vec{1}=(1,0),\,\vec{w}_{3}\to 0. (III.6)

Plugging this fixed points into the integral, we obtain the factor of proportion 222The term 1/(2)Δ11/(\infty^{2})^{\Delta_{1}} coming from w1\vec{w}_{1}\to\infty is cancelled by the Jacobian of this linear fractional transformation. as following

iλ(2π)4(i=12dyid2ziyi3)y1Δ1y2Δ2(y22+|z21|2)Δ2dω3ω3ω3Δ3y14(m12y22+m22z22)m13m22(y22+z22)2\displaystyle i\lambda(2\pi)^{4}\displaystyle{\int\left(\prod_{i=1}^{2}\frac{dy_{i}d^{2}z_{i}}{y_{i}^{3}}\right)\frac{y_{1}^{\Delta_{1}}y_{2}^{\Delta_{2}}}{(y_{2}^{2}+|\vec{z}_{2}-\vec{1}|^{2})^{\Delta_{2}}}\int\frac{d\omega_{3}}{\omega_{3}}\,\omega_{3}^{\Delta_{3}}\,\frac{y_{1}^{4}(m_{1}^{2}y_{2}^{2}+m_{2}^{2}\vec{z}_{2}^{2})}{m_{1}^{3}m_{2}^{2}(y_{2}^{2}+\vec{z}_{2}^{2})^{2}}}
×δ(y1m1m2y2(y22+z22)m12y22+m22z22)δ2(z1m22(y22+z22)z2m12y22+m22z22)δ(ω3y2(m12m22)2m2(y22+z22)).\displaystyle\times\delta(y_{1}-\frac{m_{1}m_{2}y_{2}(y_{2}^{2}+\vec{z}_{2}^{2})}{m_{1}^{2}y_{2}^{2}+m_{2}^{2}\vec{z}_{2}^{2}})\delta^{2}(\vec{z}_{1}-\frac{m_{2}^{2}(y_{2}^{2}+\vec{z}_{2}^{2})\vec{z}_{2}}{m_{1}^{2}y_{2}^{2}+m_{2}^{2}\vec{z}_{2}^{2}})\delta(\omega_{3}-\frac{y_{2}(m_{1}^{2}-m_{2}^{2})}{2m_{2}(y_{2}^{2}+\vec{z}_{2}^{2})}). (III.7)

Eliminating the delta functions we are left with a three-fold integral in a single hyperbolic space

iλ(2π)4(m12m22)Δ312Δ31m12Δ1m2Δ1+Δ3𝑑y2d2z2y2Δ1+Δ2+Δ33(y22+|z21|2)Δ2(y22+z22)Δ3Δ1(z22+m12m22y22)Δ1.\displaystyle\frac{i\lambda(2\pi)^{4}(m_{1}^{2}-m_{2}^{2})^{\Delta_{3}-1}}{2^{\Delta_{3}-1}m_{1}^{2-\Delta_{1}}m_{2}^{\Delta_{1}+\Delta_{3}}}\displaystyle{\int dy_{2}d^{2}z_{2}\frac{y_{2}^{\Delta_{1}+\Delta_{2}+\Delta_{3}-3}}{(y_{2}^{2}+|\vec{z}_{2}-\vec{1}|^{2})^{\Delta_{2}}\,(y_{2}^{2}+\vec{z}_{2}^{2})^{\Delta_{3}-\Delta_{1}}\,(\vec{z}_{2}^{2}+\frac{m_{1}^{2}}{m_{2}^{2}}y_{2}^{2})^{\Delta_{1}}}}. (III.8)

This integral can be computed using the Feynman parametrization. The denominator under the Feynman parametrization becomes

1(y22+|z21|2)Δ2(y22+z22)Δ3Δ1(z22+m12m22y22)Δ1=Γ(Δ2+Δ3)Γ(Δ2)Γ(Δ3Δ1)Γ(Δ1)\displaystyle\frac{1}{(y_{2}^{2}+|\vec{z}_{2}-\vec{1}|^{2})^{\Delta_{2}}\,(y_{2}^{2}+\vec{z}_{2}^{2})^{\Delta_{3}-\Delta_{1}}\,(\vec{z}_{2}^{2}+\frac{m_{1}^{2}}{m_{2}^{2}}y_{2}^{2})^{\Delta_{1}}}=\frac{\Gamma(\Delta_{2}+\Delta_{3})}{\Gamma(\Delta_{2})\Gamma(\Delta_{3}-\Delta_{1})\Gamma(\Delta_{1})}
×01dt101t1dt2t1Δ21t2Δ3Δ11(1t1t2)Δ11[(z2t11)2+t1(1t1)+y22(t1+t2+(1t1t2)m12m22)]Δ2+Δ3.\displaystyle\times\int_{0}^{1}dt_{1}\int_{0}^{1-t_{1}}dt_{2}\frac{t_{1}^{\Delta_{2}-1}t_{2}^{\Delta_{3}-\Delta_{1}-1}(1-t_{1}-t_{2})^{\Delta_{1}-1}}{\big{[}(\vec{z}_{2}-t_{1}\vec{1})^{2}+t_{1}(1-t_{1})+y_{2}^{2}\big{(}t_{1}+t_{2}+(1-t_{1}-t_{2})\frac{m_{1}^{2}}{m_{2}^{2}}\big{)}\big{]}^{\Delta_{2}+\Delta_{3}}}. (III.9)

Then the integration over z2\vec{z}_{2} is the standard Euclidean integral encountered in dimensional regularization

d2z2\displaystyle\int d^{2}z_{2} 1[(z2t11)2+t1(1t1)+y22(t1+t2+(1t1t2)m12m22)]Δ2+Δ3\displaystyle\frac{1}{\big{[}(\vec{z}_{2}-t_{1}\vec{1})^{2}+t_{1}(1-t_{1})+y_{2}^{2}\big{(}t_{1}+t_{2}+(1-t_{1}-t_{2})\frac{m_{1}^{2}}{m_{2}^{2}}\big{)}\big{]}^{\Delta_{2}+\Delta_{3}}}
=πΓ(Δ2+Δ31)Γ(Δ2+Δ3)1[t1(1t1)+y22(t1+t2+(1t1t2)m12m22)]Δ2+Δ31.\displaystyle=\frac{\pi\Gamma(\Delta_{2}+\Delta_{3}-1)}{\Gamma(\Delta_{2}+\Delta_{3})}\frac{1}{\big{[}t_{1}(1-t_{1})+y_{2}^{2}\big{(}t_{1}+t_{2}+(1-t_{1}-t_{2})\frac{m_{1}^{2}}{m_{2}^{2}}\big{)}\big{]}^{\Delta_{2}+\Delta_{3}-1}}. (III.10)

The integration over y2y_{2} can be computed using the Euler beta function

0𝑑y2\displaystyle\int_{0}^{\infty}dy_{2} y2Δ1+Δ2+Δ33[t1(1t1)+y22(t1+t2+(1t1t2)m12m22)]Δ2+Δ31\displaystyle\frac{y_{2}^{\Delta_{1}+\Delta_{2}+\Delta_{3}-3}}{\big{[}t_{1}(1-t_{1})+y_{2}^{2}\big{(}t_{1}+t_{2}+(1-t_{1}-t_{2})\frac{m_{1}^{2}}{m_{2}^{2}}\big{)}\big{]}^{\Delta_{2}+\Delta_{3}-1}}
=Γ(Δ1+Δ2+Δ321)Γ(Δ2+Δ3Δ12)2Γ(Δ2+Δ31)[t1(1t1)]Δ1Δ2Δ32[t1+t2+(1t1t2)m12m22]Δ1+Δ2+Δ321.\displaystyle=\frac{\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1)\Gamma(\frac{\Delta_{2}+\Delta_{3}-\Delta_{1}}{2})}{2\Gamma(\Delta_{2}+\Delta_{3}-1)}\frac{\big{[}t_{1}(1-t_{1})\big{]}^{\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}}}{\big{[}t_{1}+t_{2}+(1-t_{1}-t_{2})\frac{m_{1}^{2}}{m_{2}^{2}}\big{]}^{\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1}}. (III.11)

Finally only the integration of Feynman parameters are left

01𝑑t101t1𝑑t2t1Δ1+Δ2Δ321t2Δ3Δ11(1t1t2)Δ11(1t1)Δ1Δ2Δ32[t1+t2+(1t1t2)m12m22]Δ1+Δ2+Δ321.\displaystyle\int_{0}^{1}dt_{1}\int_{0}^{1-t_{1}}dt_{2}\frac{t_{1}^{\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}-1}t_{2}^{\Delta_{3}-\Delta_{1}-1}(1-t_{1}-t_{2})^{\Delta_{1}-1}(1-t_{1})^{\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}}}{\big{[}t_{1}+t_{2}+(1-t_{1}-t_{2})\frac{m_{1}^{2}}{m_{2}^{2}}\big{]}^{\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1}}. (III.12)

Using the famous Mellin-Barnes method, this integral can be computed into an infinite series of the mass ratio m2/m1m_{2}/m_{1} as following

Γ(Δ1)Γ(Δ3Δ1)Γ(Δ1+Δ2+Δ321)Γ(Δ1Δ2Δ32+1)Γ(Δ3)n=0{(1)nn!(m2m1)Δ1+Δ2+Δ32+2n\displaystyle\frac{\Gamma(\Delta_{1})\Gamma(\Delta_{3}-\Delta_{1})}{\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1)\Gamma(\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}+1)\Gamma(\Delta_{3})}\sum_{n=0}^{\infty}\,\Bigg{\{}\frac{(-1)^{n}}{n!}\left(\frac{m_{2}}{m_{1}}\right)^{\Delta_{1}+\Delta_{2}+\Delta_{3}-2+2n}
×Γ(1Δ2n)Γ(Δ1+Δ2Δ32+n)Γ(Δ1+Δ2+Δ321+n)\displaystyle\times\Gamma(1-\Delta_{2}-n)\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}+n)\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+n)
×F12(Δ3Δ1,Δ1+Δ2+Δ321+nΔ3;1m22m12)+(Δ2Δ~2)},\displaystyle\times{}_{2}F_{1}\left(\begin{array}[]{c}\Delta_{3}-\Delta_{1},\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+n\\ \Delta_{3}\end{array};1-\frac{m_{2}^{2}}{m_{1}^{2}}\right)+(\Delta_{2}\rightarrow\tilde{\Delta}_{2})\Bigg{\}}, (III.15)

To keep focused on the main physics, we put the computation of this series into the appendix 333However, it is hard to carry out the summation of this series into a concise and clean expression. In the appendix, we show that taking the massless limit m20m_{2}\to 0 of this infinite series is equivalent to the massless limit of this integral. The extra information obtained from this infinite series is the contribution of subleading terms, as discussed in section IV.1. and return to this complete result when discuss subleading terms of the massless limit in section IV.1.

Let’s focus on the physics here. Because this is a well-defined finite integral, when the mass m2m_{2} is decreasing, it does not touch any singularity nor branch cut. So it is safe to take the massless limit m20m_{2}\to 0 in this finite integral (III.12). Under this limit, the integral simplifies as

(m2m1)Δ1+Δ2+Δ3201𝑑t101t1𝑑t2t1Δ1+Δ2Δ321t2Δ3Δ11(1t1t2)Δ1Δ2Δ32\displaystyle\left(\frac{m_{2}}{m_{1}}\right)^{\Delta_{1}+\Delta_{2}+\Delta_{3}-2}\int_{0}^{1}dt_{1}\int_{0}^{1-t_{1}}dt_{2}\,t_{1}^{\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}-1}t_{2}^{\Delta_{3}-\Delta_{1}-1}(1-t_{1}-t_{2})^{\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}}
×(1t1)Δ1Δ2Δ32=(m2m1)Δ1+Δ2+Δ32Γ(1Δ2)Γ(Δ1+Δ2Δ32)Γ(Δ3Δ1)Γ(1+Δ3Δ1Δ22),\displaystyle{\quad}\times(1-t_{1})^{\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}}=\left(\frac{m_{2}}{m_{1}}\right)^{\Delta_{1}+\Delta_{2}+\Delta_{3}-2}\frac{\Gamma(1-\Delta_{2})\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2})\Gamma(\Delta_{3}-\Delta_{1})}{\Gamma(1+\frac{\Delta_{3}-\Delta_{1}-\Delta_{2}}{2})}, (III.16)

where only the leading term is kept because all subleading terms vanish when combining with the phase factor of  (II) and the limit Δ2=1\Delta_{2}=1, as will be shown in the next section. We see that a special gamma function Γ(1Δ2)\Gamma(1-\Delta_{2}) appears in the massless limit m20m_{2}\to 0. It has a pole in the conformal soft limit Δ21\Delta_{2}\to 1. This pole 1/(1Δ2)1/(1-\Delta_{2}) is the key result of our paper. It is crucial for the massless limit of this 3pt amplitude. Without it, the massless limit of ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle would be zero, which is unphysical because ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle exists for sure.

IV Massless limit m20m_{2}\to 0 and conformally soft Δ21\Delta_{2}\to 1

For this 3pt amplitude, the massless limit m20m_{2}\to 0 makes sense physically. The kinematics only requires the masses to satisfy m1m2m_{1}\geq m_{2}. It is legitimate to imagine situations where the outgoing particle becomes lighter and lighter. As long as m2<m1m_{2}<m_{1}, no abrupt change of physical process happens. So it is safe to imagine that when m20m_{2}\to 0, the resulting amplitude becomes ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle. To be consistent with the phase factor in the massless limit of conformal primary waves (II), the massless limit of the 3pt amplitude is

(m22)Δ2Γ(Δ2)πΓ(Δ21)ΦΔ1m1ΦΔ2m2ϕΔ3\overunderset?m20ΦΔ1m1ϕΔ2ϕΔ3\displaystyle\left(\frac{m_{2}}{2}\right)^{\Delta_{2}}\frac{\Gamma\left(\Delta_{2}\right)}{\pi\Gamma(\Delta_{2}-1)}\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle\overunderset{?}{m_{2}\to 0}{\Longrightarrow}\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle

where the question mark means this equality has to be proved. Below we show that this equality is valid only for Δ2=1\Delta_{2}=1. Both amplitudes have the same functional dependence on w\vec{w}, so we only need to check the factor of proportion. Note that here we associate the phase factor of ϕΔ2\phi_{\Delta_{2}} of  (II) and people may ask why we did not use the phase factor of \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}} of  (II). Here the leading order of the massless limit requires Δ2=1\Delta_{2}=1 and the two phase factors of ϕΔ2\phi_{\Delta_{2}} and \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}} are the same. At subleading orders of the massless limit, these two phase factors are different and details will be discussed in section IV.1.

Combining the factors in  (III.8)-(III) and (III), the left hand side of (IV) under m20m_{2}\to 0 is

iλ(2π)4m1Δ3Δ222Δ2+Δ3m222Δ2Γ(1Δ2)Γ(Δ21)Γ(Δ1+Δ2+Δ321)Γ(Δ1+Δ2Δ32)Γ(Δ2+Δ3Δ12)Γ(Δ1)Γ(1+Δ3Δ1Δ22).\displaystyle i\lambda(2\pi)^{4}\frac{m_{1}^{\Delta_{3}-\Delta_{2}-2}}{2^{\Delta_{2}+\Delta_{3}}m_{2}^{2-2\Delta_{2}}}\frac{\Gamma(1-\Delta_{2})}{\Gamma(\Delta_{2}-1)}\frac{\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1)\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2})\Gamma(\frac{\Delta_{2}+\Delta_{3}-\Delta_{1}}{2})}{\Gamma(\Delta_{1})\Gamma(1+\frac{\Delta_{3}-\Delta_{1}-\Delta_{2}}{2})}. (IV.1)

This is a well-defined finite limit only if Δ21\Delta_{2}\to 1. So the massless limit m20m_{2}\to 0 requires the conformal soft limit Δ2=1\Delta_{2}=1, if decreasing m2m_{2} is a well-behaved physical process. We see that the numerator Γ(1Δ2)\Gamma(1-\Delta_{2}) coming from the massive amplitude are crucial in the conformal soft limit. It gives the leading contribution in the combined limit. The denominator Γ(Δ21)\Gamma(\Delta_{2}-1) coming from the primary wave function can be viewed as a ’normalization’, which balances the pole of the numerator and thus making the final amplitude well-defined.

Now with both m20m_{2}\to 0 and Δ2=1\Delta_{2}=1, the factors on both sides of (IV) take the same value 444There is a total minus sign in the left hand side of (IV), which is irrelevant for the physics discussion in our paper.

iλ(2π)4m1Δ332Δ3+1Γ(Δ1+Δ312)Γ(Δ1+1Δ32)Γ(Δ1).\displaystyle i\lambda(2\pi)^{4}\frac{m_{1}^{\Delta_{3}-3}}{2^{\Delta_{3}+1}}\frac{\Gamma(\frac{\Delta_{1}+\Delta_{3}-1}{2})\Gamma(\frac{\Delta_{1}+1-\Delta_{3}}{2})}{\Gamma(\Delta_{1})}. (IV.2)

This concludes the finding of this paper

(m22)Δ2Γ(Δ2)πΓ(Δ21)ΦΔ1m1ΦΔ2m2ϕΔ3m20Δ21ΦΔ1m1ϕΔ2=1ϕΔ3.\displaystyle\left(\frac{m_{2}}{2}\right)^{\Delta_{2}}\frac{\Gamma\left(\Delta_{2}\right)}{\pi\Gamma(\Delta_{2}-1)}\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle\xRightarrow[m_{2}\to 0]{\Delta_{2}\to 1}\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}=1}\phi_{\Delta_{3}}\rangle. (IV.3)

IV.1 Subleading term m22nm_{2}^{2n} and analytically continued Δ2=1n\Delta_{2}=1-n

For convenience, we give the complete result of the 3pt amplitude here

ΦΔ1m1ΦΔ2m2ϕΔ3=iλ(2π)4(m12m22)Δ312Δ3m12Δ1m2Δ1+Δ3πΓ(Δ2+Δ3Δ12)Γ(Δ2)Γ(Δ3)Γ(Δ1Δ2Δ32+1)\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle=i\lambda(2\pi)^{4}\frac{(m_{1}^{2}-m_{2}^{2})^{\Delta_{3}-1}}{2^{\Delta_{3}}m_{1}^{2-\Delta_{1}}m_{2}^{\Delta_{1}+\Delta_{3}}}\frac{\pi\Gamma(\frac{\Delta_{2}+\Delta_{3}-\Delta_{1}}{2})}{\Gamma(\Delta_{2})\Gamma(\Delta_{3})\Gamma(\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}+1)}
×n=0{(1)nn!(m2m1)Δ1+Δ2+Δ32+2nΓ(1Δ2n)Γ(Δ1+Δ2+Δ321+n)\displaystyle\times\sum_{n=0}^{\infty}\,\Bigg{\{}\frac{(-1)^{n}}{n!}\left(\frac{m_{2}}{m_{1}}\right)^{\Delta_{1}+\Delta_{2}+\Delta_{3}-2+2n}\Gamma(1-\Delta_{2}-n)\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+n)
×Γ(Δ1+Δ2Δ32+n)F12(Δ3Δ1,Δ1+Δ2+Δ321+nΔ3;1m22m12)+(Δ2Δ~2)}\displaystyle\times\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}+n){}_{2}F_{1}\left(\begin{array}[]{c}\Delta_{3}-\Delta_{1},\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+n\\ \Delta_{3}\end{array};1-\frac{m_{2}^{2}}{m_{1}^{2}}\right)+(\Delta_{2}\rightarrow\tilde{\Delta}_{2})\Bigg{\}} (IV.6)
1Γ(Δ2)n=0{m2Δ2(22n)Γ(1nΔ2)+m2Δ2+2nΓ(Δ21n)},\displaystyle\propto\ldots\frac{1}{\Gamma(\Delta_{2})}\sum_{n=0}^{\infty}\Bigg{\{}m_{2}^{\Delta_{2}-(2-2n)}\Gamma(1-n-\Delta_{2})\ldots+m_{2}^{-\Delta_{2}+2n}\Gamma(\Delta_{2}-1-n)\ldots\Bigg{\}}, (IV.7)

where the sum is symmetric under the switch Δ2Δ~2\Delta_{2}\leftrightarrow\tilde{\Delta}_{2} and we emphasize the key quantities in the last line. Note that the total amplitude is not symmetric under the switch Δ2Δ~2\Delta_{2}\leftrightarrow\tilde{\Delta}_{2}, because it is an amplitude of ΦΔ2m2\Phi_{\Delta_{2}}^{m_{2}}, not of ΦΔ2m2+ΦΔ~2m2\Phi_{\Delta_{2}}^{m_{2}}+\Phi_{\tilde{\Delta}_{2}}^{m_{2}}. Then it is obvious that the poles of the two parts of the sum can not contribute in a ’symmetric’ way to the total amplitude, although they are symmetric inside the sum. The 1/Γ(Δ2)1/\Gamma(\Delta_{2}) plays the key role in this distinction between the poles the two parts.

Now we continue to analyze the subleading orders of the massless limit. In (A), there is an infinite number of subleading terms m22n,m_{2}^{2n}, n>0n>0, associated with functions Γ(1Δ2n)\Gamma(1-\Delta_{2}-n) and Γ(1Δ~2n)\Gamma(1-\tilde{\Delta}_{2}-n). They can also contribute poles when the scaling dimensions are analytically continued to Δ2=1n\Delta_{2}=1-n or Δ~2=1n\tilde{\Delta}_{2}=1-n. Now at subleading orders we can distinguish between the two phase factors of ϕΔ2\phi_{\Delta_{2}} and \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}} in (II). The poles coming from Γ(1Δ2n)\Gamma(1-\Delta_{2}-n) or Γ(1Δ~2n)\Gamma(1-\tilde{\Delta}_{2}-n) must be cancelled by the denominator of the phase factors Γ(Δ21)\Gamma(\Delta_{2}-1) or Γ(Δ~21)\Gamma(\tilde{\Delta}_{2}-1) respectively.

[I]: Firstly let’s consider the case of ϕΔ2\phi_{\Delta_{2}} that we used in (IV). Let’s pick the n-th term from the series (A) by analytically continued dimension Δ2=1n\Delta_{2}=1-n. The first part of the sum (terms of dimension Δ2\Delta_{2}) captures the corresponding subleading term of m2m_{2}, but the second part of the sum (terms of dimension Δ~2\tilde{\Delta}_{2}) has an extra factor m22nm_{2}^{2n} that is much more subleading. This is shown in the following

ΦΔ1m1ΦΔ2m2ϕΔ3\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle 1Γ(1n){m21+n+m21+3n}\displaystyle\propto\frac{1}{\Gamma(1-n)}\Bigg{\{}m_{2}^{-1+n}\ldots+m_{2}^{-1+3n}\ldots\Bigg{\}}
ϕΔ2=1n+m24n1Γ(1n)ϕ~Δ~2=1+n(using  (II))\displaystyle\propto\phi_{\Delta_{2}=1-n}+m_{2}^{4n}\frac{1}{\Gamma(1-n)}\tilde{\phi}_{\tilde{\Delta}_{2}=1+n}\,(\textrm{using ~{}\eqref{eq:basislimit}})
ϕΔ2=1n.\displaystyle\propto\phi_{\Delta_{2}=1-n}. (IV.8)

So the contribution of the second part (of dimension Δ~2\tilde{\Delta}_{2}) vanish at the specific subleading order of the massless limit. In detail, the first part of the sum contributes to the left hand side of (IV) as

iλ(2π)4\displaystyle i\lambda(2\pi)^{4} m1Δ3Δ222n2Δ2+Δ3m222Δ22nΓ(1Δ2n)Γ(Δ21)(1)nn!\displaystyle\frac{m_{1}^{\Delta_{3}-\Delta_{2}-2-2n}}{2^{\Delta_{2}+\Delta_{3}}m_{2}^{2-2\Delta_{2}-2n}}\frac{\Gamma(1-\Delta_{2}-n)}{\Gamma(\Delta_{2}-1)}\frac{(-1)^{n}}{n!}
×Γ(Δ1+Δ2+Δ321+n)Γ(Δ2+Δ3Δ12)Γ(Δ1+Δ2Δ32+n)Γ(Δ1Δ2Δ32+1n)Γ(Δ1)Γ(Δ1Δ2Δ32+1)Γ(1n+Δ3Δ1Δ22).\displaystyle\times\frac{\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+n)\Gamma(\frac{\Delta_{2}+\Delta_{3}-\Delta_{1}}{2})\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}+n)\Gamma(\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}+1-n)}{\Gamma(\Delta_{1})\Gamma(\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}+1)\Gamma(1-n+\frac{\Delta_{3}-\Delta_{1}-\Delta_{2}}{2})}. (IV.9)

When the scaling dimension is analytically continued to Δ2=1n\Delta_{2}=1-n, this contribution is finite and well-behaved as

iλ(2π)4m1Δ3+Δ242Δ2+Δ3Γ(Δ1+Δ2Δ32)Γ(Δ1Δ2+Δ32)Γ(Δ1).\displaystyle i\lambda(2\pi)^{4}\frac{m_{1}^{\Delta_{3}+\Delta_{2}-4}}{2^{\Delta_{2}+\Delta_{3}}}\frac{\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2})\Gamma(\frac{\Delta_{1}-\Delta_{2}+\Delta_{3}}{2})}{\Gamma(\Delta_{1})}. (IV.10)

This means that the subleading term m22nm_{2}^{2n} combined with analytically continued Δ2=1n\Delta_{2}=1-n leads to the 3pt amplitude 555For subleading terms, there is a subtlety here. The analytic continuation Δ2=1n\Delta_{2}=1-n has to be taken before the massless limit m20m_{2}\to 0, because they have preceding terms 1,m22,,m22n21,m_{2}^{2},\ldots,m_{2}^{2n-2} that are more singular in the massless limit. This subtlety is absent for the leading term.

(m22)Δ2Γ(Δ2)πΓ(Δ21)ΦΔ1m1ΦΔ2m2ϕΔ3m20Δ21nΦΔ1m1ϕΔ2=1nϕΔ3.\displaystyle\left(\frac{m_{2}}{2}\right)^{\Delta_{2}}\frac{\Gamma\left(\Delta_{2}\right)}{\pi\Gamma(\Delta_{2}-1)}\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle\xRightarrow[m_{2}\to 0]{\Delta_{2}\to 1-n}\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}=1-n}\phi_{\Delta_{3}}\rangle. (IV.11)

The nontrivial massless limit at these subleading orders does give the exact 3pt amplitude ΦΔ1m1ϕΔ2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle. We see that the analytic continuation on scaling dimensions is a powerful tool in selecting subleading contributions in the massless limit of massive amplitude. It happens that these analytically continued dimensions belong to the range of generalized conformal primary operators Δ20\Delta\in 2-\mathbb{Z}_{\geqslant 0} of massless bosons [13].

[II]: Secondly let’s consider the case of \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}} to see if the following equality works

(m22)Δ~2Γ(Δ~2)πΓ(Δ~21)ΦΔ1m1ΦΔ2m2ϕΔ3\overunderset?m20ΦΔ1m1\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTSϕΔ3\displaystyle\left(\frac{m_{2}}{2}\right)^{\tilde{\Delta}_{2}}\frac{\Gamma\left(\tilde{\Delta}_{2}\right)}{\pi\Gamma(\tilde{\Delta}_{2}-1)}\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle\overunderset{?}{m_{2}\to 0}{\Longrightarrow}\langle\Phi_{\Delta_{1}}^{m_{1}}\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}\phi_{\Delta_{3}}\rangle

Now pick the n-th term from the series (A) by analytically continued dimension Δ~2=1n\tilde{\Delta}_{2}=1-n. The first part of the sum (terms of dimension Δ2\Delta_{2}) has an extra factor m22nm_{2}^{2n} that is much more subleading, and only the second part of the sum (terms of dimension Δ~2\tilde{\Delta}_{2}) captures the corresponding subleading term. This is shown in the following

ΦΔ1m1ΦΔ2m2ϕΔ3\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle 1Γ(1+n){m21+3n+m21+n}\displaystyle\propto\frac{1}{\Gamma(1+n)}\Bigg{\{}m_{2}^{-1+3n}\ldots+m_{2}^{-1+n}\ldots\Bigg{\}}
m24nϕΔ2=1+n+1Γ(1+n)ϕ~Δ~2=1n(using  (II))\displaystyle\propto m_{2}^{4n}\phi_{\Delta_{2}=1+n}+\frac{1}{\Gamma(1+n)}\tilde{\phi}_{\tilde{\Delta}_{2}=1-n}\,(\textrm{using ~{}\eqref{eq:basislimit}})
1Γ(1+n)ϕ~Δ~2=1n.\displaystyle\propto\frac{1}{\Gamma(1+n)}\tilde{\phi}_{\tilde{\Delta}_{2}=1-n}. (IV.12)

So the contribution of the first part (of dimension Δ2\Delta_{2}) vanish at the specific subleading order of the massless limit. We can also see that the subleading terms is not exactly the massless primaries, because of the overall factor. Only for the subleading order n=2n=2 we have the massless primary because Γ(2)=1\Gamma(2)=1.

Combining it with all the factors, the second part of the sum contributes to the left hand side of (IV.1) as

(1)n+1iλ(2π)4m1Δ3+Δ~24(n1)!2Δ~2+Δ3Γ(Δ1+Δ2Δ32)Γ(Δ2+Δ3Δ12)Γ(1+Δ1+Δ2+Δ32)Γ(Δ2)Γ(Δ1)Γ(1+Δ3Δ1Δ22).\displaystyle\frac{(-1)^{n+1}i\lambda(2\pi)^{4}m_{1}^{\Delta_{3}+\tilde{\Delta}_{2}-4}}{(n-1)!2^{\tilde{\Delta}_{2}+\Delta_{3}}}\frac{\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2})\Gamma(\frac{\Delta_{2}+\Delta_{3}-\Delta_{1}}{2})\Gamma(-1+\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2})}{\Gamma(\Delta_{2})\Gamma(\Delta_{1})\Gamma(1+\frac{\Delta_{3}-\Delta_{1}-\Delta_{2}}{2})}. (IV.13)

Compared with the 3pt shadowed amplitude of (II.1), we obtain the massless limit

(m22)Δ~2Γ(Δ~2)πΓ(Δ~21)ΦΔ1m1ΦΔ2m2ϕΔ3\overundersetΔ21+nΔ~21nm20(1)n+1(n1)!Γ(Δ2=1+n)ΦΔ1m1\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2=1n\stretchto\scaleto\SavedStyle.54670.5OcFTSϕΔ3\displaystyle\left(\frac{m_{2}}{2}\right)^{\tilde{\Delta}_{2}}\frac{\Gamma\left(\tilde{\Delta}_{2}\right)}{\pi\Gamma(\tilde{\Delta}_{2}-1)}\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle\overunderset{\overset{\tilde{\Delta}_{2}\to 1-n}{\Delta_{2}\to 1+n}}{m_{2}\to 0}{\Longrightarrow}\frac{(-1)^{n+1}}{(n-1)!\Gamma(\Delta_{2}=1+n)}\langle\Phi_{\Delta_{1}}^{m_{1}}\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}=1-n}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}\phi_{\Delta_{3}}\rangle

There is a mismatch by the strange factor 1/(n1)!Γ(Δ2=1+n)1/(n-1)!\Gamma(\Delta_{2}=1+n) at the n-th subleading order. Only the first subleading order n=1n=1 gives an exact match, whose operator dimension is Δ2=2\Delta_{2}=2. This is easy to understand: the total amplitude (IV.1) is not symmetric under the switch Δ2Δ~2\Delta_{2}\leftrightarrow\tilde{\Delta}_{2}, because of the overall prefactor 1/Γ(Δ2)1/\Gamma(\Delta_{2}).

All other subleading orders with dimension Δ23\Delta_{2}\geq 3 do not give an exact match, and they happen to be out of the range of generalized conformal primary operators Δ20\Delta\in 2-\mathbb{Z}_{\geqslant 0} of massless bosons [13]. So the subleading orders of the massless limit associated with \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}} manifests the constraint of generalized conformal primary operators Δ20\Delta\in 2-\mathbb{Z}_{\geqslant 0}. Only for the generalized conformal primary operator it can reduce to the exact 3pt amplitude ΦΔ1m1\ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTSϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}}\phi_{\Delta_{3}}\rangle.

IV.2 Vanish of the 3pt celestial amplitude ϕΔ1ϕΔ2ϕΔ3\langle\phi_{\Delta_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle

We further check the massless limit and conformal soft limit. Take the 3pt amplitude

ΦΔ1m1ϕΔ2ϕΔ3(m1)Δ2+Δ34Γ(Δ1+Δ2Δ32)Γ(Δ1Δ2+Δ32)Γ(Δ1).\displaystyle\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle\propto\left(m_{1}\right)^{\Delta_{2}+\Delta_{3}-4}\frac{\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2})\Gamma(\frac{\Delta_{1}-\Delta_{2}+\Delta_{3}}{2})}{\Gamma(\Delta_{1})}. (IV.14)

This amplitude does not have a function like Γ(1Δ1)\Gamma(1-\Delta_{1}), so there is no pole to sustain the amplitude in combined limit. Take the massless limit m10m_{1}\to 0

ϕΔ1ϕΔ2ϕΔ3(m1)Δ1Γ(Δ1)πΓ(Δ11)ΦΔ1m1ϕΔ2ϕΔ3.\displaystyle\langle\phi_{\Delta_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle\propto\left(m_{1}\right)^{\Delta_{1}}\frac{\Gamma\left(\Delta_{1}\right)}{\pi\Gamma(\Delta_{1}-1)}\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle. (IV.15)

In the conformal soft limit Δ11\Delta_{1}\to 1, it is either zero or infinite

ϕΔ1ϕΔ2ϕΔ3iλ1(m1)1+iλ1+iλ2+iλ3,\displaystyle\langle\phi_{\Delta_{1}}\phi_{\Delta_{2}}\phi_{\Delta_{3}}\rangle\propto i\lambda_{1}\left(m_{1}\right)^{-1+i\lambda_{1}+i\lambda_{2}+i\lambda_{3}}, (IV.16)

depending on which one of m1m_{1} and λ1\lambda_{1} goes to zero faster. The zero or infinity mean that it does not have a limiting amplitudes of three massless scalars. This might be another explanation of the fact that 3pt celestial massless amplitudes vanish due to 4d kinematics.

V Conclusion

In this work, we studied the massless limit of the 3pt celestial amplitude of two massive states ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle. In the massless limit m20m_{2}\to 0, it reduces to the 3pt celestial amplitude of one massive state ΦΔ1m1ϕΔ2=1ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\phi_{\Delta_{2}=1}\phi_{\Delta_{3}}\rangle, provided with the conformal soft limit Δ21\Delta_{2}\to 1. The pole 1/(1Δ2)1/(1-\Delta_{2}) coming from Γ(1Δ2)\Gamma(1-\Delta_{2}) of the massless limit are crucial for the physics here. This can be compared with the conformal soft limit of celestial gluons, where the soft energy ω0\omega\to 0 gives the leading contribution and a pole of 1/(Δ1)1/(\Delta-1) arises. For celestial gluons, higher-point amplitudes are used, because the 3pt amplitude vanish due to 4d kinematics and the amplitudes reduce to lower-point amplitudes during the conformal soft limit. Here, the presence of massive states preserves the 3pt amplitude and it does not reduce to lower-point amplitude.

This connection between the m20m_{2}\to 0 and the Δ21\Delta_{2}\to 1 is unexpected and interesting, because naively one expects that the resulting amplitude under the m20m_{2}\to 0 limit can have arbitrary dimension Δ2\Delta_{2}. But the results from this celestial amplitude show that it is only consistent for the very special dimension Δ2=1\Delta_{2}=1, which is related with the conformal soft modes of massless primaries. To understand the detailed mechanism of this connection, a new and independent analysis is necessary that is different from the celestial amplitude approach, for example, an analysis from the asymptotic symmetry. This is one of the open question.

We also find that the analytic continuation of scaling dimensions is a powerful tool in selecting subleading orders of the massless limit of massive amplitudes. These scaling dimensions Δ2=1n\Delta_{2}=1-n and Δ2=2\Delta_{2}=2 fall into the range of generalized conformal primary operators Δ20\Delta\in 2-\mathbb{Z}_{\geqslant 0} of massless bosons [13], which are also the scaling dimensions of the w1+\textrm{w}_{1+\infty} algebra [17]. These generalized conformal primaries are obtained from analytic constraints on the massless primaries and the w1+\textrm{w}_{1+\infty} algebra is from the current algebra of asymptotic symmetry. It is interesting that these generalized conformal primaries are recovered from the massless limits of massive amplitude. Can we dig more from this connection? This is another open question.

It would be interesting and meaningful to know what would happen when all three particles are massive ΦΔ1m1ΦΔ2m2ΦΔ3m3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\Phi_{\Delta_{3}}^{m_{3}}\rangle. When taking the massless limit, for example m30m_{3}\to 0, would it reduce to the 3pt amplitude ΦΔ1m1ΦΔ2m2ϕΔ3\langle\Phi_{\Delta_{1}}^{m_{1}}\Phi_{\Delta_{2}}^{m_{2}}\phi_{\Delta_{3}}\rangle together with Δ31\Delta_{3}\to 1? However, this 3pt celestial amplitude of three massive scalars are hard to compute. After eliminating the momentum conservating delta function, a five-fold integral is left. In the near-extremal limit m1=2(1+ϵ)mm_{1}=2(1+\epsilon)m, m2=m3=mm_{2}=m_{3}=m, ϵ0\epsilon\to 0, the five-fold integral reduces to a three-fold integral in a single hyperbolic space, which gives the tree-level 3pt Witten diagram at the leading order of ϵ\sqrt{\epsilon} [10]. For general mass configurations, the five-fold integral of hyperbolic coordinates is very complicated. Furthermore, two of the five integration variables are coupled and have nontrival end points in their integration region depending on the masses. Currently this five-fold integral of hyperbolic coordinates remains to be an open problem.

Appendix A The integral of Feynman parameters for general masses

In this appendix, we compute the integral of Feynman parameters (III.12). Firstly we do a change of variable

t2=(1t1)s,s[0,1],\displaystyle t_{2}=(1-t_{1})s,\quad s\in[0,1], (A.1)

to decouple the integration regions of this double-integral. This leads to

01𝑑t101𝑑st1Δ1+Δ2Δ321sΔ3Δ11(1t1)Δ2(1s)Δ11[t11t1+s+(1s)m12m22]Δ1+Δ2+Δ321.\displaystyle\int_{0}^{1}dt_{1}\int_{0}^{1}ds\,\frac{t_{1}^{\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}-1}s^{\Delta_{3}-\Delta_{1}-1}(1-t_{1})^{-\Delta_{2}}(1-s)^{\Delta_{1}-1}}{\big{[}\frac{t_{1}}{1-t_{1}}+s+(1-s)\frac{m_{1}^{2}}{m_{2}^{2}}\big{]}^{\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1}}. (A.2)

Then the denominator can be rewritten using the famous Mellin-Barnes formula to decouple the integrand of this double-integral

1[t11t1+s+(1s)m12m22]Δ1+Δ2+Δ321=\displaystyle\frac{1}{\big{[}\frac{t_{1}}{1-t_{1}}+s+(1-s)\frac{m_{1}^{2}}{m_{2}^{2}}\big{]}^{\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1}}= 1Γ(Δ1+Δ2+Δ321)iidα2πi(t1/(1t1))α[s+(1s)m12m22]Δ1+Δ2+Δ321+α\displaystyle\frac{1}{\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1)}\displaystyle{\int_{-i\infty}^{i\infty}\frac{d\alpha}{2\pi i}}\frac{(t_{1}/(1-t_{1}))^{\alpha}}{\big{[}s+(1-s)\frac{m_{1}^{2}}{m_{2}^{2}}\big{]}^{\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+\alpha}}
×Γ(Δ1+Δ2+Δ321+α)Γ(α),\displaystyle\times{}\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+\alpha)\Gamma(-\alpha), (A.3)

with the integration contour to be 1/2<Re(α)<0-1/2<\operatorname{Re}(\alpha)<0.

Now the two integrals of t1t_{1} and ss can be done independently. The t1t_{1} integration gives

01𝑑t1t1Δ1+Δ2Δ321+α(1t1)Δ2α=Γ(1Δ2α)Γ(Δ1+Δ2Δ32+α)Γ(Δ1Δ2Δ32+1).\displaystyle\int_{0}^{1}dt_{1}\,\frac{t_{1}^{\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}-1+\alpha}}{(1-t_{1})^{-\Delta_{2}-\alpha}}=\frac{\Gamma(1-\Delta_{2}-\alpha)\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}+\alpha)}{\Gamma(\frac{\Delta 1-\Delta 2-\Delta 3}{2}+1)}. (A.4)

And the ss integration gives

01𝑑s\displaystyle\int_{0}^{1}ds\, sΔ3Δ11(1s)Δ11[s+(1s)m12m22]Δ1+Δ2+Δ321+α=Γ(Δ1)Γ(Δ3Δ1)Γ(Δ3)\displaystyle\frac{s^{\Delta_{3}-\Delta_{1}-1}(1-s)^{\Delta_{1}-1}}{\big{[}s+(1-s)\frac{m_{1}^{2}}{m_{2}^{2}}\big{]}^{\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+\alpha}}=\frac{\Gamma(\Delta_{1})\Gamma(\Delta_{3}-\Delta_{1})}{\Gamma(\Delta_{3})}
×(m2m1)Δ1+Δ2+Δ32+2αF12(Δ3Δ1,Δ1+Δ2+Δ321+αΔ3;1m22m12).\displaystyle\times\left(\frac{m_{2}}{m_{1}}\right)^{\Delta_{1}+\Delta_{2}+\Delta_{3}-2+2\alpha}{}_{2}F_{1}\left(\begin{array}[]{c}\Delta_{3}-\Delta_{1},\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+\alpha\\ \Delta_{3}\end{array};1-\frac{m_{2}^{2}}{m_{1}^{2}}\right). (A.7)

Finally we perform the Mellin-Barnes integral of α\alpha. Because the scaling dimensions are on the principal continuous series Δ=1+iλ,λ\Delta=1+i\lambda,\lambda\in\mathbb{R}, the poles in the integrand are well separated from each other, so there is no pinched singularity to worry about. We can safely close the contour to the right hand side of the complex plane of α\alpha, which picks the poles of Γ(α)\Gamma(-\alpha) and Γ(1Δ2α)\Gamma(1-\Delta_{2}-\alpha). The contribution of Γ(α)\Gamma(-\alpha) and Γ(1Δ2α)\Gamma(1-\Delta_{2}-\alpha) generates terms of dimension Δ2\Delta_{2} and Δ~2=2Δ2\tilde{\Delta}_{2}=2-\Delta_{2} respectively, and these terms turn out to be the same under the switch of dimension Δ2Δ~2\Delta_{2}\leftrightarrow\tilde{\Delta}_{2}. So the contribution of Γ(α)\Gamma(-\alpha) and Γ(1Δ2α)\Gamma(1-\Delta_{2}-\alpha) are from the fields ϕΔ2\phi_{\Delta_{2}} and \ThisStyle\stackengine.1\LMpt\SavedStyleϕΔ~2\stretchto\scaleto\SavedStyle.54670.5OcFTS\ThisStyle{\stackengine{-.1\LMpt}{$\SavedStyle\phi_{\tilde{\Delta}_{2}}$}{\stretchto{\scaleto{\SavedStyle\mkern 0.2mu\sim}{.5467}}{0.5}}{O}{c}{F}{T}{S}} respectively. This leads to the final result

Γ(Δ1)Γ(Δ3Δ1)Γ(Δ1+Δ2+Δ321)Γ(Δ1Δ2Δ32+1)Γ(Δ3)n=0{(1)nn!(m2m1)Δ1+Δ2+Δ32+2n\displaystyle\frac{\Gamma(\Delta_{1})\Gamma(\Delta_{3}-\Delta_{1})}{\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1)\Gamma(\frac{\Delta_{1}-\Delta_{2}-\Delta_{3}}{2}+1)\Gamma(\Delta_{3})}\sum_{n=0}^{\infty}\,\Bigg{\{}\frac{(-1)^{n}}{n!}\left(\frac{m_{2}}{m_{1}}\right)^{\Delta_{1}+\Delta_{2}+\Delta_{3}-2+2n}
×Γ(1Δ2n)Γ(Δ1+Δ2Δ32+n)Γ(Δ1+Δ2+Δ321+n)\displaystyle\times\Gamma(1-\Delta_{2}-n)\Gamma(\frac{\Delta_{1}+\Delta_{2}-\Delta_{3}}{2}+n)\Gamma(\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+n)
×F12(Δ3Δ1,Δ1+Δ2+Δ321+nΔ3;1m22m12)+(Δ2Δ~2)},\displaystyle\times{}_{2}F_{1}\left(\begin{array}[]{c}\Delta_{3}-\Delta_{1},\frac{\Delta_{1}+\Delta_{2}+\Delta_{3}}{2}-1+n\\ \Delta_{3}\end{array};1-\frac{m_{2}^{2}}{m_{1}^{2}}\right)+(\Delta_{2}\rightarrow\tilde{\Delta}_{2})\Bigg{\}}, (A.10)

where the second part of the sum is obtained from the first part by replacing Δ2\Delta_{2} with Δ~2\tilde{\Delta}_{2}. At the leading order of the massless limit m20m_{2}\to 0, only the n=0n=0 term contributes and these two parts are the same under the conformal soft limit Δ2=Δ~2=1\Delta_{2}=\tilde{\Delta}_{2}=1. When the scaling dimension is analytically continued to Δ2=1n\Delta_{2}=1-n or Δ~2=1n\tilde{\Delta}_{2}=1-n respectively, subleading orders of the massless limit will be picked up, which is discussed in section IV.1.

Acknowledgements.
Wei Fan is supported in part by the National Natural Science Foundation of China under Grant No. 12105121.

References