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

Coupled-channel scattering on a torus

Peng Guo pguo@jlab.org Thomas Jefferson National Accelerator Facility, Newport News, VA 23606, USA    Jozef J. Dudek Thomas Jefferson National Accelerator Facility, Newport News, VA 23606, USA Department of Physics, Old Dominion University, Norfolk, VA 23529, USA    Robert G. Edwards Thomas Jefferson National Accelerator Facility, Newport News, VA 23606, USA    Adam P. Szczepaniak Physics Department, Indiana University, Bloomington, IN 47405, USA Center For Exploration of Energy and Matter, Indiana University, Bloomington, IN 47408, USA.
(September 28, 2025)
Abstract

Based on the Hamiltonian formalism approach, a generalized Lüscher’s formula for two particle scattering in both the elastic and coupled-channel cases in moving frames is derived from a relativistic Lippmann-Schwinger equation. Some strategies for extracting scattering amplitudes for a coupled-channel system from the discrete finite-volume spectrum are discussed and illustrated with a toy model of two-channel resonant scattering. This formalism will, in the near future, be used to extract information about hadron scattering from lattice QCD computations.

pacs:
12.38.Gc, 11.80.Gw, 11.80.Jy, 13.75.Lb
preprint: JLAB-THY-12-1659

I Introduction

Hadron spectroscopy in lattice QCD is entering a new era, in particular, recent developments in the application of variational methods Michael:1985ne ; Luscher:1990ck ; Blossier:2009kd to large bases of hadron interpolating fields have made the extraction of the excited spectrum of hadronic states a realistic possibility (see e.g. Jo:2010 ; Edwards:2011 ; Jo_scatt:2012 ). Since excited hadrons appear as resonances in the continuous distribution of multi-hadron scattering states, to study hadron spectroscopy one requires evaluation of scattering amplitudes, but because lattice QCD is formulated in Euclidean space, we do not have direct access to these Maiani:1990ca . Fortunately, in a finite volume, interactions between particles as they evolve from the in to the out states lead to discrete changes in a free particle’s energy that can be related to the scattering amplitude Lusher:1991 .

Various extensions to the framework derived by Lüscher in Lusher:1991 have been proposed which allow for evaluation outside the center-of-mass frame Gottlieb:1995 ; Lin:2001 ; Christ:2005 ; Bernard:2007 ; Bernard:2008 , and to include the coupled-channel effects that can appear above the inelastic threshold Liu:2005 ; Doring:2011 ; Aoki:2011 ; Briceno:2012yi ; Hansen:2012tf . The original approach and its extensions to describe the moving center-of-mass frame have been quite successfully used by the lattice community to extract elastic hadron-hadron scattering phase shiftsAoki:2007rd ; Sasaki:2008 ; Feng:2011 ; Jo_scatt:2011 ; Jo_scatt:2012 ; Beane_scatt:2012 ; Lang_scatt:2011 ; Aoki:2011yj .

In this work, we discuss a generalization of Lüscher’s method for relativistic scattering in terms of a Hamiltonian where the specific interactions considered are based on a relativistic particle exchange model. We apply the technique to a two-channel system and a generalized Lüscher’s equation for scattering in a moving frame is derived based on the relativistic Lippmann-Schwinger equation. The coupled-channel system has been considered previously, Liu:2005 ; Doring:2011 ; Briceno:2012yi ; Hansen:2012tf , and our result agrees with these works. A novelty of the present work is to discuss practical strategies for extraction of scattering amplitude parameters from lattice simulations of a coupled-channel system. These strategies are demonstrated using an explicit toy model of resonant two-channel scattering.

The paper is organized as follows. A discussion of elastic scattering in a finite-volume is given in Section II, with extension to the coupled channel system in Section III. Strategies for extracting scattering amplitudes from measured discrete finite-volume spectra are presented in Section IV. The summary and outlook are given in Section V.

II Finite-volume elastic scattering in a Hamiltonian framework

In this section we present relativistic two-particle scattering on a torus using the Hamiltonian formalism developed in Glazek:1993 ; Bakker:2003 . In particular we consider a complex scalar field, Φ\Phi, describing a charged boson, ϕ±\phi^{\pm}, of mass mm, and its interactions with a neutral boson, θ\theta, which acts as a force carrier and is described by a real scalar field, Θ\Theta. We first derive the Lüscher formula describing the finite-volume spectrum of the asymptotic two-particle, ϕ+ϕ\phi^{+}\phi^{-} state,

det[δJM,JMcotδJ(k)JM,JM(𝐐)(k)]=0,\det\left[\delta_{JM,J^{\prime}M^{\prime}}\cot\delta_{J}(k)-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k)\right]=0,

where the volume and scattering-frame dependent matrix element JM,JM(𝐐)\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}} is defined in Eq.(B) and (B), and the center-of-mass frame scattering momentum, kk, is related to the energy by s/4m2\sqrt{s/4-m^{2}}. The model corresponds to a Lagrangian density,

=μΦμΦm2ΦΦ+12μΘμΘ12μ2Θ2gΘΦΦ,\mathcal{L}=\partial_{\mu}\Phi^{*}\partial^{\mu}\Phi-m^{2}\Phi^{*}\Phi+\tfrac{1}{2}\partial_{\mu}\Theta\partial^{\mu}\Theta-\tfrac{1}{2}\mu^{2}\Theta^{2}-g\Theta\Phi^{*}\Phi, (1)

from which the Hamiltonian can be derived following the canonical procedure of instant-time quantization (see Appendix A) Dirac:1949 . Taking matrix elements of the Hamiltonian in an infinite basis of Fock states spanned by any number of ϕ\phi and θ\theta bosons one can obtain a Schrödinger equation H^|Ψ=E|Ψ\hat{H}|\Psi\rangle=E|\Psi\rangle for the eigenstates of the theory. Assuming μm\mu\gg m, in describing low-energy ϕ\phi-boson scattering we can truncate the Fock space to include up to one θ\theta-boson in the intermediate state, which reduces the Schrödinger equation to

[H22H23H32H33][|ϕ+ϕ|ϕ+ϕθ]\displaystyle\begin{bmatrix}H_{22}&H_{23}\\ H_{32}&H_{33}\end{bmatrix}\begin{bmatrix}|\phi^{+}\phi^{-}\rangle\\ |\phi^{+}\phi^{-}\theta\rangle\end{bmatrix} =\displaystyle= E[|ϕ+ϕ|ϕ+ϕθ].\displaystyle E\begin{bmatrix}|\phi^{+}\phi^{-}\rangle\\ |\phi^{+}\phi^{-}\theta\rangle\end{bmatrix}. (2)

The three-particle sector can be formally eliminated, resulting in an effective two-body equation,

(EH22)|ϕ+ϕ=H231EH33H32|ϕ+ϕ.(E-H_{22})|\phi^{+}\phi^{-}\rangle=H_{23}\frac{1}{E-H_{33}}H_{32}|\phi^{+}\phi^{-}\rangle. (3)

II.1 Two-particle scattering in an infinite volume

Before considering two-particle states on a torus, we will first review the scattering problem in infinite-volume, with further details given in Appendix A. After eliminating the three-particle states |ϕ+ϕθ|\phi^{+}\phi^{-}\theta\rangle from the coupled system (cf Eq.(2)) we are left with an equation for the center-of-mass frame momentum-space wavefunction, φJM(𝐪)\varphi_{JM}(\mathbf{q}), which is a product of a radial wavefunction depending on the magnitude of the relative 3-momentum, q=|𝐪|q=|{\bf q}|, and the spherical harmonic of definite angular momentum, (J,M)(J,M),

φJM(𝐪)=1s2𝐪2+m2d3𝐤(2π)3V(𝐪,𝐤)φJM(𝐤).\varphi_{JM}(\mathbf{q})=\frac{1}{\sqrt{s}-2\sqrt{\mathbf{q}^{2}+m^{2}}}\int\!\!\frac{d^{3}\mathbf{k}}{(2\pi)^{3}}\,V(\mathbf{q},\mathbf{k})\,\varphi_{JM}(\mathbf{k}). (4)

Here, E=sE=\sqrt{s} is the energy of the two-particle system in the center-of-mass frame. The non-local potential, V(𝐪,𝐤)V(\mathbf{q},\mathbf{k}), induced by θ\theta-exchange is given explicitly in Eq. (26). Expressing this equation in coordinate space via a Fourier transform gives

ψJM(𝐫)=d3𝐫G0(𝐫𝐫;s)d3𝐳V~(𝐫,𝐳)ψJM(𝐳),\psi_{JM}(\mathbf{r})=\int\!\!d^{3}\mathbf{r}^{\prime}\,G_{0}(\mathbf{r}-\mathbf{r}^{\prime};\sqrt{s})\int\!\!d^{3}\mathbf{z}\,\tilde{V}(\mathbf{r}^{\prime},-\mathbf{z})\,\psi_{JM}(\mathbf{z}), (5)

where the free Green’s function is given by

G0(𝐫𝐫;s)=d3𝐪(2π)3ei𝐪(𝐫𝐫)s2𝐪2m2.G_{0}(\mathbf{r}-\mathbf{r}^{\prime};\sqrt{s})=\int\!\!\frac{d^{3}\mathbf{q}}{(2\pi)^{3}}\,\frac{e^{i\mathbf{q}\cdot(\mathbf{r}-\mathbf{r}^{\prime})}}{\sqrt{s}-2\sqrt{\mathbf{q}^{2}-m^{2}}}. (6)

The wavefunction satisfies a relativistic Schrödinger equation,

(s22+m2)ψJM(𝐫)=d3𝐳V~(𝐫,𝐳)ψJM(𝐳).\left(\sqrt{s}-2\sqrt{-\nabla^{2}+m^{2}}\right)\psi_{JM}(\mathbf{r})=\int\!\!d^{3}\mathbf{z}\,\tilde{V}(\mathbf{r},-\mathbf{z})\,\psi_{JM}(\mathbf{z}). (7)

While Eq.(5) was derived in the context of a particular model, our subsequent derivation only requires the general form of the relativistic Lippmann-Schwinger equation. The asymptotic component of the two-body wavefunction relevant to scattering is given by the large distance behavior of the Green’s function. Evaluating the integral in Eq.(6) (cf. Appendix A), we find

G0\displaystyle G_{0} (𝐫;s=2k2+m2)=\displaystyle(\mathbf{r};\sqrt{s}=2\sqrt{k^{2}+m^{2}})=
s2eikr4πr1rmρdρ(2π)2ρ2m2eρrk2+ρ2,\displaystyle-\frac{\sqrt{s}}{2}\,\frac{e^{ikr}}{4\pi r}-\frac{1}{r}\int_{m}^{\infty}\!\!\frac{\rho\,d\rho}{(2\pi)^{2}}\,\sqrt{\rho^{2}-m^{2}}\frac{e^{-\rho r}}{k^{2}+\rho^{2}}, (8)

with the first term on the right hand side dominating as rr\to\infty. For a potential V~\tilde{V} which falls at large separations, the solution to Eq (7) outside the range of the interaction is given by

ψJM(𝐫)\displaystyle\psi_{JM}(\mathbf{r}) s2miJYJM(𝐫^)[4πjJ(kr)+ikfJ(k)hJ+(kr)],\displaystyle\to\frac{\sqrt{s}}{2m}i^{J}\,Y_{JM}(\hat{\mathbf{r}})\left[4\pi\,j_{J}(kr)+ik\,f_{J}(k)\,h^{+}_{J}(kr)\right], (9)

where fJ(k)f_{J}(k) is the partial wave scattering amplitude,

fJ(k)=miJd3𝐫d3𝐳jJ(kr)YJM(𝐫^)V~(𝐫,𝐳)ψJM(𝐳),f_{J}(k)=-\frac{m}{i^{J}}\!\int\!\!d^{3}\mathbf{r}^{\prime}d^{3}\mathbf{z}\,j_{J}(kr^{\prime})\,Y^{*}_{JM}(\hat{\mathbf{r}}^{\prime})\,\tilde{V}(\mathbf{r}^{\prime},-\mathbf{z})\,\psi_{JM}(\mathbf{z}), (10)

which up to the inelastic threshold can be parameterized in terms of a single real momentum-dependent parameter, the scattering phase-shift, δJ(k)\delta_{J}(k), as

fJ(k)=4πkeiδJsinδJ.f_{J}(k)=\frac{4\pi}{k}e^{i\delta_{J}}\sin\delta_{J}.

II.2 Two-particle scattering on a torus

Now we consider the theory in a cubic box of volume V=L3V=L^{3}, with periodic boundary conditions. In Eq. (5) we split the integral over 𝐫\mathbf{r}^{\prime} into a sum of integrals over a set of boxes labelled by the integers 𝐧\mathbf{n} representing the location of one of its corners,

ψJM(L)(𝐫)\displaystyle\psi^{(L)}_{JM}(\mathbf{r}) =𝐧3L3d3𝐫G0(𝐫𝐫𝐧L;s)\displaystyle=\sum_{\mathbf{n}\in\mathbb{Z}^{3}}\int_{L^{3}}\!\!d^{3}\mathbf{r}^{\prime}\,G_{0}(\mathbf{r}-\mathbf{r}^{\prime}-\mathbf{n}L;\sqrt{s})
×d3𝐳V~(𝐫+𝐧L,𝐳𝐧L)ψJM(L)(𝐳+𝐧L).\displaystyle\times\int\!\!d^{3}\mathbf{z}^{\prime}\,\tilde{V}(\mathbf{r}^{\prime}+\mathbf{n}L,-\mathbf{z}^{\prime}-\mathbf{n}L)\,\psi^{(L)}_{JM}(\mathbf{z}^{\prime}+\mathbf{n}L). (11)

In general we can make the wavefunctions periodic up to a phase,

ψJM(L)(𝐳+𝐧L)=ei𝐐𝐧LψJM(L)(𝐳),\psi^{(L)}_{JM}(\mathbf{z}+\mathbf{n}L)=e^{i\mathbf{Q}\cdot\mathbf{n}L}\psi^{(L)}_{JM}(\mathbf{z}),

where the Bloch wave-vector, 𝐐\mathbf{Q}, is related to the total momentum of the two-particle system Gottlieb:1995 by 𝐏=2γ𝐐\mathbf{P}=2\gamma\,\mathbf{Q}. γ=s+𝐏2/s\gamma=\sqrt{s+\mathbf{P}^{2}}/\sqrt{s} is the Lorentz contraction factor that reduces the effective size of the box in the direction parallel to 𝐏\mathbf{P}. Using the periodicity of the potential, V~(𝐫+𝐧L,𝐳𝐧L)=V~(𝐫,𝐳)\tilde{V}(\mathbf{r}^{\prime}+\mathbf{n}L,-\mathbf{z}^{\prime}-\mathbf{n}L)=\tilde{V}(\mathbf{r}^{\prime},-\mathbf{z}^{\prime}), and the boundary condition on the wavefunction, we have

ψJM(L,𝐐)(𝐫)\displaystyle\psi^{(L,\mathbf{Q})}_{JM}(\mathbf{r}) =L3d3𝐫G𝐐(𝐫𝐫;s)\displaystyle=\int_{L^{3}}\!\!d^{3}\mathbf{r}^{\prime}\,G_{\mathbf{Q}}(\mathbf{r}-\mathbf{r}^{\prime};\sqrt{s})
×d3𝐳V~(𝐫,𝐳)ψJM(L,𝐐)(𝐳),\displaystyle\quad\quad\times\int\!\!d^{3}\mathbf{z}\,\tilde{V}(\mathbf{r}^{\prime},-\mathbf{z})\,\psi^{(L,\mathbf{Q})}_{JM}(\mathbf{z}),

which is analogous to the infinite-volume equation, but with the Green’s function given by

G𝐐(𝐫𝐫;s)=𝐧3G0(𝐫𝐫𝐧L;s)ei𝐐𝐧L.G_{\mathbf{Q}}(\mathbf{r}-\mathbf{r}^{\prime};\sqrt{s})=\sum_{\mathbf{n}\in\mathbb{Z}^{3}}G_{0}(\mathbf{r}-\mathbf{r}^{\prime}-\mathbf{n}L;\sqrt{s})\,e^{i\mathbf{Q}\cdot\mathbf{n}L}.

Using the Poisson summation formula, (2π)3𝐧3ei𝐐𝐧L=L3𝐧3δ(𝐐+2πL𝐧)(2\pi)^{-3}\sum_{\mathbf{n}\in\mathbb{Z}^{3}}e^{i\mathbf{Q}\cdot\mathbf{n}L}=L^{-3}\sum_{\mathbf{n}\in\mathbb{Z}^{3}}\delta\left(\mathbf{Q}+\tfrac{2\pi}{L}\mathbf{n}\right), we obtain

G𝐐(𝐫𝐫;s)\displaystyle G_{\mathbf{Q}}(\mathbf{r}-\mathbf{r}^{\prime};\sqrt{s}) =1L3𝐪P𝐐ei𝐪(𝐫𝐫)s2𝐪2+m2\displaystyle=\frac{1}{L^{3}}\sum_{\mathbf{q}\in P_{\mathbf{Q}}}\frac{e^{i\mathbf{q}\cdot(\mathbf{r}-\mathbf{r}^{\prime})}}{\sqrt{s}-2\sqrt{\mathbf{q}^{2}+m^{2}}}
s21L3𝐪P𝐐ei𝐪(𝐫𝐫)k2𝐪2,\displaystyle\to\frac{\sqrt{s}}{2}\frac{1}{L^{3}}\sum_{\mathbf{q}\in P_{\mathbf{Q}}}\frac{e^{i\mathbf{q}\cdot(\mathbf{r}-\mathbf{r}^{\prime})}}{k^{2}-\mathbf{q}^{2}},

where P𝐐={𝐪3|𝐪=2πL𝐧+𝐐,for𝐧3}P_{\mathbf{Q}}=\{\mathbf{q}\in\mathbb{R}^{3}|\,\mathbf{q}=\tfrac{2\pi}{L}\mathbf{n}+\mathbf{Q},\;\text{for}\;\mathbf{n}\in\mathbb{Z}^{3}\}, and where we have retained only the leading term in the limit L|𝐫𝐫|m1L\gg|{\bf r}-{\bf r^{\prime}}|\gg m^{-1}. Finally, expanding Eq.(11) for rrr\gg r^{\prime} and using the definition of the scattering amplitude, Eq. (10) we can express the wavefunction as

ψJM(L,𝐐)(𝐫)\displaystyle\psi^{(L,\mathbf{Q})}_{JM}(\mathbf{r}) s2m(k)iJfJ(k)JMYJM(𝐫^)\displaystyle\to\frac{\sqrt{s}}{2m}(-k)\,i^{J}f_{J}(k)\sum_{J^{\prime}M^{\prime}}Y_{J^{\prime}M^{\prime}}(\hat{\mathbf{r}})
×[δJM,JMnJ(kr)JM,JM(𝐐)(k)jJ(kr)].\displaystyle\times\left[\delta_{JM,J^{\prime}M^{\prime}}\,n_{J^{\prime}}(kr)-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k)\;j_{J^{\prime}}(kr)\right]. (12)

The residual sum over all angular momenta reflects the broken rotational invariance induced by the finite cubic volume, with the volume-dependent matrix elements {\cal M} given in Appendix B. In the infinite-volume case, the most general solution of the relativistic Schrödinger equation, Eq.(7), outside the range of the potential is JMcJMψJM(𝐫)\sum_{JM}c_{JM}\,\psi_{JM}(\mathbf{r}) for ψJM(𝐫)\psi_{JM}(\mathbf{r}) given by Eq. (9). Correspondingly in finite-volume, the most general solution is given by JMcJMψJM(L,𝐐)(𝐫)\sum_{JM}c_{JM}\,\psi^{(L,\mathbf{Q})}_{JM}(\mathbf{r}) for ψJM(L,𝐐)(𝐫)\psi^{(L,\mathbf{Q})}_{JM}(\mathbf{r}) given by Eq.(12). Matching the two wavefunctions at a fixed rr, larger than the range of the interaction, we obtain

JMcJMYJM(𝐫^)iJ[4πjJ(kr)+ikfJ(kr)hJ+(kr)]\displaystyle\sum_{JM}c_{JM}Y_{JM}(\hat{\bf r})\,i^{J}\left[4\pi\,j_{J}(kr)+ik\,f_{J}(kr)\,h^{+}_{J}(kr)\right]
=JM,JMcJMiJkfJ(k)YJM(𝐫^)\displaystyle\quad=-\!\!\!\sum_{JM,J^{\prime}M^{\prime}}\!\!c_{JM}\,i^{J}k\,f_{J}(k)\,Y_{J^{\prime}M^{\prime}}(\hat{\bf r})
×[δJM,JMnJ(kr)JM,JM(𝐐)(k)jJ(kr)],\displaystyle\quad\quad\times\left[\delta_{JM,J^{\prime}M^{\prime}}\,n_{J^{\prime}}(kr)-{\cal M}^{(\bf Q)}_{JM,J^{\prime}M^{\prime}}(k)\,j_{J^{\prime}}(kr)\right],

which has a non-trivial, cJM0c_{JM}\neq 0, solution provided

det[δJM,JMcotδJ(k)JM,JM(𝐐)(k)]=0.\det\left[\delta_{JM,J^{\prime}M^{\prime}}\,\cot\delta_{J}(k)-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k)\right]=0. (13)

This condition expresses the relationship between the asymptotic behavior of the two-particle wavefunction on a torus, expressed through the matrix elements JM,JM(𝐐)\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}, and the effect of the interaction on the wavefunction determined by the phase shifts, δJ\delta_{J}. In practice for a given set of elastic phase-shifts, δJ(k)\delta_{J}(k), it determines a discrete spectrum of states in a finite volume.

The analysis presented here can be generalized to an arbitrary shaped box. In general the three edges of box are spanned by three arbitrary vectors 𝐋1,2,3\mathbf{L}_{1,2,3}. The volume of the cube L3L^{3} is replaced by (𝐋1×𝐋2)𝐋3(\mathbf{L}_{1}\times\mathbf{L}_{2})\cdot\mathbf{L}_{3} and the vector 𝐧L\mathbf{n}L by i=1,2,3ni𝐋i\sum_{i=1,2,3}n_{i}\mathbf{L}_{i}, nin_{i}\in\mathbb{Z}. Finally the momentum 𝐪=2π𝐧/L,𝐧3\mathbf{q}=2\pi\mathbf{n}/L,\mathbf{n}\in\mathbb{Z}^{3} is replaced by generalized momentum 2πi=1,2,3ni(𝐋j×𝐋k)/|(𝐋1×𝐋2)𝐋3|,ni2\pi\sum_{i=1,2,3}n_{i}(\mathbf{L}_{j}\times\mathbf{L}_{k})/|(\mathbf{L}_{1}\times\mathbf{L}_{2})\cdot\mathbf{L}_{3}|,\,n_{i}\in\mathbb{Z} , where indices (i,j,k)(i,j,k) follow the cyclic permutation. Such a generalization has to be considered when using the moving center-of-mass frame since the symmetric shape of a cubic box in the rest frame is deformed due to Lorentz contraction Gottlieb:1995 . In this case if 𝐏=2π𝐝/L,𝐝3\mathbf{P}=2\pi\mathbf{d}/L,\mathbf{d}\in\mathbb{Z}^{3}, is the center-of-mass momentum, the volume of the box becomes γL3\gamma L^{3}, and the vectors 𝐧L\mathbf{n}L and 2π𝐧/L2\pi\mathbf{n}/L are replaced by γ𝐧L\gamma\mathbf{n}L and 2πγ1𝐧/L2\pi\gamma^{-1}\mathbf{n}/L, respectively (using the notation defined in Gottlieb:1995 ). With these substitutions and the relation 𝐏=2γ𝐐\mathbf{P}=2\gamma\mathbf{Q}, our definition of the matrix elements JM,JM(𝐐)\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}} becomes identical to the matrix elements Mlm,lm𝐝M^{\mathbf{d}}_{lm,l^{\prime}m^{\prime}} in Eq.(89) of Gottlieb:1995 .

Typically, as discussed in Lusher:1991 , for the low-energy region that we are interested in, higher partial waves become progressively smaller and can be ignored, so that the partial wave basis can be truncated at a certain maximal angular momentum JmaxJ_{\mathrm{max}}. For a finite-volume with cubic boundaries, the continuous rotation symmetry is reduced to the little-group of allowed cubic rotations that leave the centre-of-mass momentum invariant - the matrices in Eq.(13) become block-diagonal if subduced according to the irreducible representations of these little groups. Details of subduction in general moving frames can be found in Christopher_subduction:2012 .

III Coupled channel scattering in finite volume

We extend the model of the previous section to include additional two-particle asymptotic states, by adding another species of charged bosons, σ±\sigma^{\pm}, which also couples to the force carrier, θ\theta, into the Lagrangian. We can obtain coupled equations for the two-particle states, |ϕ+ϕ|\phi^{+}\phi^{-}\rangle and |σ+σ|\sigma^{+}\sigma^{-}\rangle, by eliminating states featuring three or more particles and obtain a two-channel Schrödinger equation,

|ϕ+ϕ\displaystyle\big{|}\phi^{+}\phi^{-}\big{\rangle} =1EHϕ(0)[Vϕϕ|ϕ+ϕ+Vϕσ|σ+σ],\displaystyle=\frac{1}{E-H_{\phi}^{(0)}}\left[V_{\phi\phi}\big{|}\phi^{+}\phi^{-}\big{\rangle}+V_{\phi\sigma}\big{|}\sigma^{+}\sigma^{-}\big{\rangle}\right],
|σ+σ\displaystyle\big{|}\sigma^{+}\sigma^{-}\big{\rangle} =1EHσ(0)[Vσϕ|ϕ+ϕ+Vσσ|σ+σ],\displaystyle=\frac{1}{E-H_{\sigma}^{(0)}}\left[V_{\sigma\phi}\big{|}\phi^{+}\phi^{-}\big{\rangle}+V_{\sigma\sigma}\big{|}\sigma^{+}\sigma^{-}\big{\rangle}\right], (14)

where Hϕ(0)H_{\phi}^{(0)}, Hσ(0)H_{\sigma}^{(0)} are the one-particle operators and Vϕϕ,Vϕσ,Vσϕ,VσσV_{\phi\phi},V_{\phi\sigma},V_{\sigma\phi},V_{\sigma\sigma} are effective interactions (two-body operators) generated by the reduction to the two-particle subspace. From Eq.(14), for the channel wavefunctions, ψJMα=ϕ,σ(𝐫)𝐫|α,JM\psi^{\alpha=\phi,\sigma}_{JM}(\mathbf{r})\equiv\langle\mathbf{r}|\alpha,JM\rangle, we obtain

ψJMα(𝐫)\displaystyle\psi^{\alpha}_{JM}(\mathbf{r}) =d3𝐫G0α(𝐫𝐫;s)\displaystyle=\int\!\!d^{3}\mathbf{r}^{\prime}\,G^{\alpha}_{0}(\mathbf{r}-\mathbf{r}^{\prime};\sqrt{s})
×βd3𝐳V~αβ(𝐫,𝐳)ψJMβ(𝐳).\displaystyle\quad\times\sum_{\beta}\int\!\!d^{3}\mathbf{z}\,\tilde{V}_{\alpha\beta}(\mathbf{r}^{\prime},-\mathbf{z})\,\psi^{\beta}_{JM}(\mathbf{z}).

The coupled-channel scattering amplitudes can be defined by

fJαβ(s)=mαiJd3𝐫d3𝐳\displaystyle f_{J}^{\alpha\beta}(s)=-\frac{m_{\alpha}}{i^{J}}\int\!\!d^{3}\mathbf{r}^{\prime}d^{3}\mathbf{z}\; jJ(kαr)YJM(𝐫^)\displaystyle j_{J}(k_{\alpha}r^{\prime})\,Y_{JM}^{*}(\hat{\mathbf{r}}^{\prime})
×V~αβ(𝐫,𝐳)ψJMβ(𝐳),\displaystyle\times\tilde{V}_{\alpha\beta}(\mathbf{r}^{\prime},-\mathbf{z})\,\psi^{\beta}_{JM}(\mathbf{z}), (15)

where kα=s/4mα2k_{\alpha}=\sqrt{s/4-m_{\alpha}^{2}} is the magnitude of the relative momentum in the center-of-mass frame of the two particles in channel α\alpha. By analogy to the single channel case, the asymptotic wavefunction in channel α\alpha is given by

ψJMα(𝐫)\displaystyle\psi^{\alpha}_{JM}(\mathbf{r}) s2mαYJM(𝐫^)iJ\displaystyle\to\frac{\sqrt{s}}{2m_{\alpha}}Y_{JM}(\hat{\mathbf{r}})\,i^{J}
×[ 4πjJ(kαr)+ikαhJ+(kαr)βfJαβ(s)].\displaystyle\times\left[\,4\pi\,j_{J}(k_{\alpha}r)+ik_{\alpha}\,h^{+}_{J}(k_{\alpha}r)\,{\textstyle\sum_{\beta}}f^{\alpha\beta}_{J}(s)\right]. (16)

Extending the one-channel analysis of the asymptotic states in finite-volume to the two-channel system, one obtains,

ψJMα(L,𝐐)(𝐫)s2mα(-kα)iJJMYJM(𝐫^)[βfJαβ(s)]\displaystyle\,\psi^{\alpha(L,\mathbf{Q})}_{JM}(\mathbf{r})\to\frac{\sqrt{s}}{2m_{\alpha}}(\text{-}k_{\alpha})\,i^{J}\sum_{J^{\prime}M^{\prime}}Y_{J^{\prime}M^{\prime}}(\hat{\mathbf{r}})\left[\textstyle{\sum_{\beta}f^{\alpha\beta}_{J}(s)}\right]
×[δJM,JMnJ(kαr)JM,JM(𝐐)(kα)jJ(kαr)].\displaystyle\quad\quad\times\left[\delta_{JM,J^{\prime}M^{\prime}}\,n_{J^{\prime}}(k_{\alpha}r)-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k_{\alpha})\,j_{J^{\prime}}(k_{\alpha}r)\right].

Matching the wavefunctions in finite-volume, Eq. (LABEL:box_coupled), to the wavefunctions in infinite volume, Eq. (16), we get a determinant condition

det[δJM,JMδα,β4πkα1fJαα+[iδJM,JMJM,JM(𝐐)(kα)]fJαβfJαα]=0.\displaystyle\det\left[\delta_{JM,J^{\prime}M^{\prime}}\,\delta_{\alpha,\beta}\frac{4\pi}{k_{\alpha}}\frac{1}{f^{\alpha\alpha}_{J}}+\left[\,i\,\delta_{JM,J^{\prime}M^{\prime}}\,-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k_{\alpha})\,\right]\,\frac{f^{\alpha\beta}_{J}}{f^{\alpha\alpha}_{J}}\,\right]=0. (18)

The derivation logically extends to any number of scattering channels. Expressing the scattering amplitudes using tt-matrix elements, tαβ(J)(s)s8πfJαβ(s)t_{\alpha\beta}^{(J)}(s)\equiv\tfrac{\sqrt{s}}{8\pi}f^{\alpha\beta}_{J}(s), and introducing the phase-space for channel α\alpha by ρα(s)=2kαs\rho_{\alpha}(s)=\tfrac{2k_{\alpha}}{\sqrt{s}}, we can write the condition as

0=det[δJM,JMδα,β+iρα(s)tαβ(J)(s)(δJM,JM+iJM,JM(𝐐)(kα))],0=\det\left[\delta_{JM,J^{\prime}M^{\prime}}\delta_{\alpha,\beta}+i\rho_{\alpha}(s)\,t^{(J)}_{\alpha\beta}(s)\Big{(}\delta_{JM,J^{\prime}M^{\prime}}+i\mathcal{M}^{(\bf{Q})}_{JM,J^{\prime}M^{\prime}}(k_{\alpha})\Big{)}\right], (19)

or, alternatively in a form which expresses the effect of unitarity more directly as,

0=det[δJM,JM([t(J)(s)]αβ1+iρα(s)δα,β)δα,βρα(s)JM,JM(𝐐)(kα))].0=\det\left[\delta_{JM,J^{\prime}M^{\prime}}\Big{(}\big{[}t^{(J)}(s)\big{]}^{-1}_{\alpha\beta}+i\rho_{\alpha}(s)\delta_{\alpha,\beta}\Big{)}-\delta_{\alpha,\beta}\,\rho_{\alpha}(s)\,\mathcal{M}^{(\bf{Q})}_{JM,J^{\prime}M^{\prime}}(k_{\alpha})\Big{)}\right]. (20)

The multichannel unitarity condition can be expressed as Im[t(J)(s)]αβ1=δαβρα(s)Θ(ssthr.(α))\mathrm{Im}\big{[}t^{(J)}(s)\big{]}^{-1}_{\alpha\beta}=-\delta_{\alpha\beta}\rho_{\alpha}(s)\Theta(s-s_{\mathrm{thr.}}^{(\alpha)}), and thus, since ρα(s)\rho_{\alpha}(s) becomes pure imaginary below threshold, the first term in Eqn 20 is always real.

The form presented in Eqn 19 can be shown to be equivalent to that presented in Briceno:2012yi and Hansen:2012tf . Their expressions include an additional phase of iJJi^{J-J^{\prime}} in front of \mathcal{M}, but the effect of this phase is always cancelled in the determinant.

Reflecting the remaining symmetry of a cube in flight, there is in fact one determinant condition for each irreducible representation of the symmetry group. As presented in Jo_scatt:2012 , Christopher_subduction:2012 these conditions can be obtained by subduction, the result being conditions

0=det[δJJδnn([t(J)(s)]αβ1+iρα(s)δα,β)δα,βρα(s)Jn,Jn(𝐐,Λ)(kα))],0=\det\left[\delta_{JJ^{\prime}}\delta_{nn^{\prime}}\Big{(}\big{[}t^{(J)}(s)\big{]}^{-1}_{\alpha\beta}+i\rho_{\alpha}(s)\delta_{\alpha,\beta}\Big{)}-\delta_{\alpha,\beta}\,\rho_{\alpha}(s)\,\mathcal{M}^{(\mathbf{Q},\Lambda)}_{Jn,J^{\prime}n^{\prime}}(k_{\alpha})\Big{)}\right], (21)

where the Λ\Lambda-irrep subduced Jn,Jn(𝐐,Λ)(kα)\mathcal{M}^{(\mathbf{Q},\Lambda)}_{Jn,J^{\prime}n^{\prime}}(k_{\alpha}) is as defined in Eqn (28) of Jo_scatt:2012 . The angular-momentum space is defined by the various embeddings, nn, of spin-JJ into the irrep Λ\Lambda. If the subduction conventions in Jo_scatt:2012 are followed, for unitary tt-matrices, the conditions are purely real at all real energies.

The two-channel scattering matrix fαβf^{\alpha\beta} can be conventionally parameterizated in terms of two scattering phase-shifts, δJα(s)\delta^{\alpha}_{J}(s) (α=ϕ,σ\alpha=\phi,\sigma), and an inelasticity, ηJ(s)\eta_{J}(s), representing the fraction of flux exchanged between the two channels,

fJαα(s)=4πkαηJe2iδJα12i;fJαβ(s)=4πkαkβ1ηJ2ei(δJα+δJβ)2,\displaystyle f^{\alpha\alpha}_{J}(s)=\frac{4\pi}{k_{\alpha}}\cdot\frac{\eta_{J}\,e^{2i\delta^{\alpha}_{J}}-1}{2i};\quad\quad f^{\alpha\beta}_{J}(s)=\frac{4\pi}{\sqrt{k_{\alpha}k_{\beta}}}\cdot\frac{\sqrt{1-\eta_{J}^{2}}\,e^{i(\delta^{\alpha}_{J}+\delta^{\beta}_{J})}}{2}, (22)

so that the determinant condition can be written

det[δJM,JMcotΔJϕJM,JM(𝐐)(kϕ)kϕkσ[iδJM,JMJM,JM(𝐐)(kϕ)]1ηJ2eiΔJσ2ηJsinΔJϕkσkϕ[iδJM,JMJM,JM(𝐐)(kσ)]1ηJ2eiΔJϕ2ηJsinΔJσδJM,JMcotΔJσJM,JM(𝐐)(kσ)]=0,\displaystyle\det\begin{bmatrix}\delta_{JM,J^{\prime}M^{\prime}}\,\cot\Delta^{\phi}_{J}-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k_{\phi})&\sqrt{\tfrac{k_{\phi}}{k_{\sigma}}}\left[i\,\delta_{JM,J^{\prime}M^{\prime}}-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k_{\phi})\right]\frac{\sqrt{1-\eta_{J}^{2}}\,e^{i\Delta_{J}^{\sigma}}}{2\eta_{J}\sin\Delta_{J}^{\phi}}\\ \sqrt{\tfrac{k_{\sigma}}{k_{\phi}}}\left[i\,\delta_{JM,J^{\prime}M^{\prime}}-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k_{\sigma})\right]\frac{\sqrt{1-\eta_{J}^{2}}\,e^{i\Delta_{J}^{\phi}}}{2\eta_{J}\sin\Delta_{J}^{\sigma}}&\delta_{JM,J^{\prime}M^{\prime}}\,\cot\Delta^{\sigma}_{J}-\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k_{\sigma})\end{bmatrix}=0, (23)

where ΔJα(s)δJα(s)i2logηJ(s)\Delta_{J}^{\alpha}(s)\equiv\delta^{\alpha}_{J}(s)-\tfrac{i}{2}\log\eta_{J}(s). One can show that this result is equivalent to Eq.(4.14) in Liu:2005 .

The determinant conditions presented above for coupled-channels scattering provide only one equation, at each finite-volume energy, for many unknowns. At low energies we may be justified in only considering the effect of the lowest contributing partial-wave, but even then there are multiple unknowns. For example in the case of two-channel scattering in SS-wave we require three parameters to describe the tt-matrix at each energy which might be the two phase-shifts and inelasticity, δ0ϕ,δ0σ,η0\delta^{\phi}_{0},\,\delta^{\sigma}_{0},\,\eta_{0}. Hence additional constraints need to be imposed to obtain a unique solution. As an example, in Doring:2011 unitarized chiral perturbation was used to constrain amplitude parameters at low energies. In the next section we explore some strategies for extraction of the scattering amplitude from finite-volume spectra in the context of an analytical parametrization of the amplitude.

It is worth noting here that in a finite volume, kinematically closed channels can play a role in determining the spectrum. Examining Eqn 19, we see that the behavior of JM,JM(𝐐)(k)\mathcal{M}^{(\mathbf{Q})}_{JM,J^{\prime}M^{\prime}}(k), analytically continued below threshold (to imaginary kk) must be considered. These matrix elements decay rapidly below threshold, such that the effect of a kinematically closed channel on the finite-volume spectrum is only felt in a limited energy region below the kinematic threshold. For example if k=iκk=i\kappa,

00,00(𝐐)(iκ)=iiκ𝐧3𝐧0eκ|γ𝐧L||γ𝐧L|ei𝐐γ𝐧L,\mathcal{M}^{(\mathbf{Q})}_{00,00}(i\kappa)=i-\frac{i}{\kappa}\sum_{\begin{subarray}{c}\mathbf{n}\in\mathbb{Z}^{3}\\ {\mathbf{n}\neq 0}\end{subarray}}\frac{e^{-\kappa|\gamma\mathbf{n}L|}}{|\gamma\mathbf{n}L|}e^{i\mathbf{Q}\cdot\gamma\mathbf{n}L},

so that far below threshold, or in very large volume, i\mathcal{M}\to i, and thus in Eqn 19, the effect of this channel is removed. However this closed channel will have an effect in a region just below the threshold, for γκL1\gamma\kappa\lesssim L^{-1}. Hence we are required in practice to analytically continue scattering amplitudes below thresholds in order to determine the finite-volume spectrum. An example of this is presented in Torres:2012 for the case of SS-wave πΣ,K¯N\pi\Sigma,\overline{K}N scattering.

IV A toy model of resonant coupled-channel scattering in finite-volume

In order to explore possible strategies for extracting coupled-channel scattering amplitudes from the discrete finite-volume spectra emerging from lattice QCD computations, we consider a simple model of two-channel SS-wave scattering. The model is based on resonance-dominated scattering and satisfies multi-channel unitarity and the analytical properties required of partial wave amplitudes. With an explicit model for scattering amplitudes we can solve Eq. (19) to obtain finite-volume spectra of states as a function of the volume (V=L3V=L^{3}) and total momentum of the center-of-mass 𝐏=2π𝐝/L,𝐝3\mathbf{P}=2\pi\mathbf{d}/L,\;\mathbf{d}\in\mathbb{Z}^{3} (𝐏=2γ𝐐\mathbf{P}=2\gamma{\bf Q}). We then use this spectrum as pseudo-data representing a hypothetical lattice QCD simulation and attempt to reproduce the input model.

IV.1 Analytic model of two-channel scattering

We consider a model in which a single SS-wave resonance coupled to both scattering channels interferes with a non-resonant background. The two-channel scattering amplitude is parametrized in terms of a KK-matrix,

Kαβ(s)=gαgβM2s+γαβ(0)+γαβ(1)s+,K_{\alpha\beta}(s)=\frac{g_{\alpha}g_{\beta}}{M^{2}-s}+\gamma^{(0)}_{\alpha\beta}+\gamma^{(1)}_{\alpha\beta}\,s+\ldots, (24)

which is related to the tt-matrix by

[t1(s)]αβ=[K1(s)]αβ+δαβIα(s),\left[t^{-1}(s)\right]_{\alpha\beta}=\left[K^{-1}(s)\right]_{\alpha\beta}+\delta_{\alpha\beta}\,I_{\alpha}(s),

and to the scattering amplitude defined in Eq. (15) by tJαβ(s)=sfJαβ(s)/8πt^{\alpha\beta}_{J}(s)=\sqrt{s}f^{\alpha\beta}_{J}(s)/8\pi with (α=ϕ,σ\alpha=\phi,\sigma). Here Iα(s)I_{\alpha}(s) is the Chew-Mandelstam form Basdevant:1977 ,

Iα(s)=Iα(0)sπ4mα2𝑑s14mα2s1(ss)s,I_{\alpha}(s)=I_{\alpha}(0)-\frac{s}{\pi}\int_{4m_{\alpha}^{2}}^{\infty}\!\!\!\!ds^{\prime}\sqrt{1-\frac{4m_{\alpha}^{2}}{s^{\prime}}}\frac{1}{(s^{\prime}-s)s^{\prime}},

whose imaginary part above threshold, i.e. for s>4mα2s>4m_{\alpha}^{2}, is the negative of the phase-space, Im[Iα(s)]=ρα(s)\mathrm{Im}\big{[}I_{\alpha}(s)\big{]}=-\rho_{\alpha}(s). This form ensures the unitarity of the amplitude and provides a smooth transition across the kinematic threshold. We have opted to subtract the integral once, and it is convenient to choose Iα(0)I_{\alpha}(0) such that ReIα(M2)=0\text{Re}\,I_{\alpha}(M^{2})=0 so that we have an amplitude which for real ss near M2M^{2} is close to the Breit-Wigner form with mass MM. The tt-matrix thus constructed is an analytical function in the complex ss-plane with the discontinuity across the right-hand cut determined by unitarity.

Refer to caption
Figure 1: Phase-shifts and inelasticity for the model defined in the text.

With the following choice of parameters,

mϕ=0.2GeV,mσ=0.4GeV\displaystyle m_{\phi}=0.2\,\mathrm{GeV},\;\;m_{\sigma}=0.4\,\mathrm{GeV}
M=1.1GeV,gϕ=0.35GeV,gσ=0.2GeV\displaystyle M=1.1\,\mathrm{GeV},\;g_{\phi}=0.35\,\mathrm{GeV},\;g_{\sigma}=0.2\,\mathrm{GeV}
γϕϕ(n)=γϕσ(n)=0,γσσ(0)=0.7,γσσ(1)=0.7GeV2,γσσ(n>1)=0,\displaystyle\gamma^{(n)}_{\phi\phi}=\gamma^{(n)}_{\phi\sigma}=0,\;\gamma^{(0)}_{\sigma\sigma}=0.7,\;\gamma^{(1)}_{\sigma\sigma}=0.7\,\mathrm{GeV}^{-2},\;\gamma^{(n>1)}_{\sigma\sigma}=0,

we obtain the phase-shifts and inelasticity shown in Fig. 1. The parameters have been chosen in such a way that there is a narrow resonance near s=1.1GeV\sqrt{s}=1.1\,\mathrm{GeV}. It is usual to analyse scattering in terms of the most relevant singularities of the tt-matrix on the nearby unphysical sheets. Poles on unphysical sheets are often identified with hadron resonances. In this case the four sheets (𝗌𝗁𝖾𝖾𝗍𝖨\mathsf{sheet\,I} is the physical sheet) can be defined by

[t𝗌𝗁𝖾𝖾𝗍1(s)]αβ={[t𝖨1(s)]αβ𝗌𝗁𝖾𝖾𝗍𝖨[t𝖨1(s)]αβ+2i14mϕ2sδαϕ𝗌𝗁𝖾𝖾𝗍𝖨𝖨[t𝖨1(s)]αβ+2i14mα2sδαβ𝗌𝗁𝖾𝖾𝗍𝖨𝖨𝖨[t𝖨1(s)]αβ+2i14mσ2sδασ𝗌𝗁𝖾𝖾𝗍𝖨𝖵.\left[t_{\mathsf{sheet}}^{-1}(s)\right]_{\alpha\beta}=\begin{cases}\left[t_{\mathsf{I}}^{-1}(s)\right]_{\alpha\beta}&\mathsf{sheet\,I}\\ \left[t_{\mathsf{I}}^{-1}(s)\right]_{\alpha\beta}+2i\sqrt{1-\tfrac{4m_{\phi}^{2}}{s}}\delta_{\alpha\phi}&\mathsf{sheet\,II}\\ \left[t_{\mathsf{I}}^{-1}(s)\right]_{\alpha\beta}+2i\sqrt{1-\tfrac{4m_{\alpha}^{2}}{s}}\delta_{\alpha\beta}&\mathsf{sheet\,III}\\ \left[t_{\mathsf{I}}^{-1}(s)\right]_{\alpha\beta}+2i\sqrt{1-\tfrac{4m_{\sigma}^{2}}{s}}\delta_{\alpha\sigma}&\mathsf{sheet\,IV}\end{cases}.

The model amplitude has a single pole on the lower half-plane111and a conjugate pole on the upper half-plane of each of sheets 𝖨𝖨\mathsf{II} and 𝖨𝖨𝖨\mathsf{III}, with the tt-matrix in the neighbourhood of the pole at s0s_{0} behaving like,

[t𝗌𝗁𝖾𝖾𝗍(ss0)]αβcαcβs0s,\left[t_{\mathsf{sheet}}(s\to s_{0})\right]_{\alpha\beta}\to\frac{c_{\alpha}c_{\beta}}{s_{0}-s},

with

s0=(1.1067i20.0961)GeV\displaystyle\;\sqrt{s_{0}}=\big{(}1.1067-\tfrac{i}{2}0.0961\big{)}\,\mathrm{GeV}
cϕ=(0.3585GeV)ei0.0023π,cσ=(0.1367GeV)ei0.237π\displaystyle\;c_{\phi}=\big{(}0.3585\,\mathrm{GeV}\big{)}\,e^{-i0.0023\pi},\;c_{\sigma}=\big{(}0.1367\,\mathrm{GeV}\big{)}\,e^{-i0.237\pi}

on 𝗌𝗁𝖾𝖾𝗍𝖨𝖨\mathsf{sheet\,II} and

s0=(1.1088i20.1195)GeV\displaystyle\;\sqrt{s_{0}}=\big{(}1.1088-\tfrac{i}{2}0.1195\big{)}\,\mathrm{GeV}
cϕ=(0.3573GeV)e+i0.0026π,cσ=(0.1391GeV)e+i0.297π\displaystyle\;c_{\phi}=\big{(}0.3573\,\mathrm{GeV}\big{)}\,e^{+i0.0026\pi},\;c_{\sigma}=\big{(}0.1391\,\mathrm{GeV}\big{)}\,e^{+i0.297\pi}

on 𝗌𝗁𝖾𝖾𝗍𝖨𝖨𝖨\mathsf{sheet\,III}. Our aim is to use the finite-volume spectrum determined on a set of volumes and total momenta, 𝐏\mathbf{P}, to reproduce the pole positions of this scattering amplitude.

IV.2 Finite-volume spectrum

The finite-volume spectrum corresponding to the model defined in the previous section can be obtained by solving Eq. (23) (or equivalently Eq. (19)), where for E=s<2mσE=\sqrt{s}<2m_{\sigma} we require the analytic continuation of tt-matrix elements featuring channel σ\sigma. Assuming that partial waves higher than SS-wave are negligible reduces Eq. (23), for E>2mσE>2m_{\sigma}, to,

0\displaystyle 0 =Ω(δϕ(E),δσ(E),η(E);L,𝐝;E)\displaystyle=\Omega(\delta^{\phi}(E),\delta^{\sigma}(E),\eta(E);L,\mathbf{d};E)
=η[ϕσ]sin(δϕδσ)\displaystyle=\eta\left[\mathcal{M}_{\phi}-\mathcal{M}_{\sigma}\right]\sin\left(\delta^{\phi}-\delta^{\sigma}\right)
+[ϕ+σ]sin(δϕ+δσ)\displaystyle\quad\quad+\left[\mathcal{M}_{\phi}+\mathcal{M}_{\sigma}\right]\sin\left(\delta^{\phi}+\delta^{\sigma}\right)
η[1+ϕσ]cos(δϕδσ)\displaystyle\quad\quad-\eta\left[1+\mathcal{M}_{\phi}\,\mathcal{M}_{\sigma}\right]\cos\left(\delta^{\phi}-\delta^{\sigma}\right)
[1ϕσ]cos(δϕ+δσ),\displaystyle\quad\quad-\left[1-\mathcal{M}_{\phi}\,\mathcal{M}_{\sigma}\right]\cos\left(\delta^{\phi}+\delta^{\sigma}\right), (25)

where ϕJ=0M=0,J=0M=0(𝐐)(kϕ)\mathcal{M}_{\phi}\equiv\mathcal{M}^{(\mathbf{Q})}_{J=0\,M=0,J^{\prime}=0\,M^{\prime}=0}(k_{\phi}) with a similar expression for σ\mathcal{M}_{\sigma}. 𝐐\mathbf{Q} is a function of 𝐝\mathbf{d}, LL, EE as discussed in Sec. II.2. In Fig.2 we show the finite-volume spectrum222this would be the spectrum in irrep A1A_{1} Jo_scatt:2012 obtained by solving the determinant condition as a function of the volume in a region L=1624 GeV1L=16-24\mbox{ GeV}^{-1} (or L=3.24.7 fmL=3.2-4.7\mbox{ fm}).

Refer to caption
Figure 2: Finite-volume spectra for the KK-matrix model described in Sec. IV.1. Black dots indicate the spectrum obtained by solving Eq. (19). Red and blue curves represent the energy of a non-interacting pair of mesons (α=ϕ,σ(\alpha=\phi,\sigma) [(mα2+𝐤12+mα2+𝐤22)2𝐏2]1/2\left[\left(\sqrt{m_{\alpha}^{2}+\mathbf{k}^{2}_{1}}+\sqrt{m_{\alpha}^{2}+\mathbf{k}^{2}_{2}}\right)^{2}-\mathbf{P}^{2}\right]^{1/2}, 𝐤1+𝐤2=𝐏\mathbf{k}_{1}+\mathbf{k}_{2}=\mathbf{P} and 𝐤=2πL𝐧,𝐧3\mathbf{k}=\frac{2\pi}{L}\mathbf{n},\mathbf{n}\in\mathbb{Z}^{3}. The points labelled by letters are used as described in the text.

IV.3 “Pointwise” estimation of scattering from finite-volume spectrum

Refer to caption
Figure 3: “Pointwise” determination of the phase-shifts and inelasticity. δϕ\delta^{\phi} points below the opening of the σσ\sigma\sigma threshold (in pink) are determined by solving the elastic relation, Eqn. (13), ignoring the effect of the closed channel. The badly discrepant points just below the threshold are all within a momentum scale of 1/L1/L of the threshold where the finite-volume effect of the closed-channel should not be neglected. Energies 𝖠𝖩\mathsf{A}-\mathsf{J} are determined from constrained three-level analysis, energies 𝖺𝗁\mathsf{a}-\mathsf{h} by interpolation in δϕ\delta^{\phi} from two-level analysis. The light-colored curves show the exact input model.

In the region below 2mσ2m_{\sigma} the ϕϕ\phi\phi scattering is elastic and it is tempting to use Eq (13), completely ignoring the existence of the kinematically closed σσ\sigma\sigma channel. Doing so leads to the points shown in pink in Fig 3 which successfully reproduce the input model except in a region just below the threshold. As previously discussed, in a finite-volume the effect of a closed channel can be felt within a limited energy region immediately below the threshold. With this in mind, we must be careful not to use the simple elastic relation, Eq (13), close to a threshold.

One approach to determining the phase-shifts and inelasticities at discrete values of energy above 2mσ2m_{\sigma} is to locate multiple energy levels (in different volumes and/or different 𝐝\mathbf{d}) that appear at approximately the same energy. As an example consider the three levels labeled 𝖠\mathsf{A} in Fig. 2 which all lie within 3 MeV of s𝖠=830MeV\sqrt{s_{\mathsf{A}}}=830\,\mathrm{MeV}. For the three levels we can build three independent copies of Eq. (23) which each feature approximately the same values of δϕ(s𝖠),δσ(s𝖠),η(s𝖠)\delta^{\phi}(s_{\mathsf{A}}),\delta^{\sigma}(s_{\mathsf{A}}),\eta(s_{\mathsf{A}}), which can be determined by solving the set of simultaneous equations. Since the energies are not exactly degenerate, there need not be an exact solution to the equations and hence we seek to find the solution which minimizes

E(L,𝐝)|Ω(δϕ,δσ,η;L,𝐝;E)|2,\displaystyle\sum_{E(L,\mathbf{d})}\left|\Omega(\delta^{\phi},\delta^{\sigma},\eta;L,\mathbf{d};E)\right|^{2},

with Ω\Omega defined in Eq. (25) and where the sum is over the three energy levels. For the levels 𝖠\mathsf{A}, the obtained solution, as shown in Fig. 3, is within 3%3\% of the exact value of δϕ(s𝖠),δσ(s𝖠),η(s𝖠)\delta^{\phi}(s_{\mathsf{A}}),\delta^{\sigma}(s_{\mathsf{A}}),\eta(s_{\mathsf{A}}). We emphasize that this procedure is not guaranteed to successfully reproduce the true scattering amplitudes - the \mathcal{M} matrices can in some circumstances vary rather rapidly over a narrow energy region.

Within the energy region considered, E=0.81.2GeVE=0.8-1.2\,\mathrm{GeV}, considering only three volumes, L=16, 20, 24GeV1L=16,\,20,\,24\,\mathrm{GeV}^{-1}, and three sets of center-of-mass momentum, 𝐝=(000),(100),(110)\mathbf{d}=(000),\,(100),\,(110), we can isolate a number of such sets. These sets of three near-degenerate energy levels are labelled 𝖠𝖩\mathsf{A}-\mathsf{J} in Fig. 2. In Fig. 3 the labels are shown on the plot of δϕ\delta^{\phi} with the corresponding solutions for δσ\delta^{\sigma} and η\eta marked by solid dots.

With these points alone, in Fig. 3 we see strong hints of a signal for resonant behavior in the δϕ\delta^{\phi} phase shift. While obviously reasonably successful, this approach does not make optimal use of the finite-volume spectral information, by failing to use any energy level which does not have two near-degenerate partners. To use somewhat more of the discrete levels we might consider building a system of two instances of Eq. (23) with one parameter from the set δϕ,δσ,η\delta^{\phi},\delta^{\sigma},\eta, estimated using interpolation between already determined values. For example consider the two levels near 1.009 GeV1.009\mbox{ GeV} labeled 𝖾\mathsf{e} in Fig. 2. Linear interpolation between the energies of the 𝖢\mathsf{C} and 𝖣\mathsf{D} points in Fig. 3 gives δϕ(1.009GeV)=30.7\delta^{\phi}(1.009\,\mathrm{GeV})=30.7^{\circ}. Using this value and minimizing with respect to δσ,η\delta^{\sigma},\eta at the energy corresponding to point 𝖾\mathsf{e} we obtain δσ=50.6,η=0.903\delta^{\sigma}=50.6^{\circ},\eta=0.903. A spline interpolation using all the points 𝖠𝖩\mathsf{A-J} yields δϕ(1.009GeV)=29.6\delta^{\phi}(1.009\,\mathrm{GeV})=29.6^{\circ} which results in δσ=48.1,η=0.934\delta^{\sigma}=48.1^{\circ},\eta=0.934 at the point 𝖾\mathsf{e}. In Fig. 3 we show the results for sets of two degenerate levels labeled 𝖺𝗁\mathsf{a}-\mathsf{h} from Fig. 2. In each case the range shown indicates limiting values obtained using two methods of interpolation. Even though in some cases there is a considerable sensitivity to the interpolation method, overall the points are in reasonable agreement with the model input (solid light-colored curves).

IV.4 Parameterized estimation of scattering from finite-volume spectrum

The previously discussed “pointwise” strategy, while having the advantage of being largely model-independent, is reliant upon there being multiple energy levels which, through accident or design, are close to degenerate. Since it would be unusual to engineer lattice volumes purely for this purpose, and unusual in contemporary calculations to have such a high density of determined energy levels, it is appropriate to consider alternative methods of analysis. One such approach that makes full use of all determined levels, and which may require far fewer levels to be determined, involves parameterising the scattering amplitude and performing a minimisation to get the best description of the determined finite-volume spectrum by varying the parameters. In the current toy-model, even limited “pointwise” analysis would suggest that the phase δϕ\delta^{\phi} is rapidly rising and would indicate that a resonance could be present. In practical calculations (e.g.Jo:2010 , Dudek:2012xn ), the presence of a sharp meson resonance can also be indicated by large overlap onto fermion bilinear operators. By including a pole (as well as polynomial behaviour) in a KK-matrix parameterisation we are likely to get rapid convergence to a solution with a pole in the tt-matrix corresponding to the resonance.

We will take this opportunity to make the toy model a slightly more realistic simulation of an actual lattice QCD calculation by introducing statistical uncertainty on the energy level values. In recent work Jo_scatt:2011 , Jo_scatt:2012 , Dudek:2012xn , the Hadron Spectrum Collaboration has obtained statistical errors on excited levels as small as 0.3% and we will assume that this remains practical. For each energy level below 1.21.2 GeV on a single volume L=16GeV13.2fmL=16\,\mathrm{GeV}^{-1}\sim 3.2\,\mathrm{fm} with 𝐝=(000),(100),(110)\mathbf{d}=(000),\,(100),\,(110), we randomly generated an ensemble by drawing from a distribution whose mean is the exact value given in Fig. 2 and whose variance is chosen such that the ensemble has variance on the mean of 0.3% of the mean value. The resulting spectrum is shown in Fig. 4.

Refer to caption
Figure 4: L=16GeV1L=16\,\mathrm{GeV}^{-1}. Orange rectangles: finite-volume spectra with 0.3% noise. Black lines: exact finite-volume spectrum given in Fig.2. Also shown the position of the thresholds and the KK-matrix pole mass.
Refer to caption
Figure 5: Best-fit phase-shifts and inelasticity for the parameterizations 𝖠𝖤\mathsf{A-E} as described in the text. Solid curves with error bands show the minimized solutions with the dashed curves showing the exact input. Black dots under the energy scale show the positions of the discrete energy levels used in the minimization. Bottom-right panel shows the determined position of a tt-matrix pole on the unphysical sheets 𝖨𝖨\mathsf{II}(red) and 𝖨𝖨𝖨\mathsf{III}(blue) for the five parameterisations along with the exact input positions (black dots).

Parametrizing according to a form like that given in Eq. (24), we can minimise a function,

χ2({ai})=E𝔫(L,𝐝)[E𝔫(L,𝐝)E𝔫det(L,𝐝;{ai})]2σ(E𝔫(L,𝐝))2,\chi^{2}\big{(}\{a_{i}\}\big{)}=\sum_{E_{\mathfrak{n}}(L,\mathbf{d})}\frac{\left[E_{\mathfrak{n}}(L,\mathbf{d})-E^{\text{det}}_{\mathfrak{n}}(L,\mathbf{d};\{a_{i}\})\right]^{2}}{\sigma(E_{\mathfrak{n}}(L,\mathbf{d}))^{2}},

(where EdetE^{\text{det}} are the solutions of Eq. (19)), by varying the parameters of the KK-matrix parameterisation, {ai}={M,gϕ,gσ,γ(n)}\{a_{i}\}=\{M,g_{\phi},g_{\sigma},\gamma^{(n)}\ldots\}. In practical lattice QCD calculations, the χ2\chi^{2} can be trivially redefined to deal with correlated data by replacing the inverse diagonal variance (1/σ21/\sigma^{2}) by the inverse of the data covariance matrix.

In Fig.5 we show the parameterized phase-shifts and inelasticity obtained using five model parameterisations:

  • 𝖠\mathsf{A}: “exact model”, which uses a 1st order polynomial in ss to describe the non-pole contribution to the KK-matrix with γϕϕ(0,1)=γϕσ(0,1)0\gamma^{(0,1)}_{\phi\phi}=\gamma^{(0,1)}_{\phi\sigma}\equiv 0 [5 parameters]

  • 𝖡\mathsf{B}: “relaxed model”, with a 1st order polynomial in all channels i.e. all γ(0,1)\gamma^{(0,1)} free [9 parameters]

  • 𝖢\mathsf{C}: “tight model”, with 0th order of polynomials in all channels i.e. γ(1)=0\gamma^{(1)}=0 [6 parameters].

  • 𝖣\mathsf{D}: “loose model”, with 2nd order of polynomials in all channels i.e. γ(0,1,2)\gamma^{(0,1,2)} free [12 parameters].

  • 𝖤\mathsf{E}: “two pole model”, with no polynomial, but with a second KK-matrix pole with independent variable couplings [6 parameters].

As one would expect, within statistical uncertainty, parameterization 𝖠\mathsf{A} reproduces the input model. Parameterisation 𝖡\mathsf{B}, which is more flexible, also reproduces the input quite well over the energy region where data is given, albeit with a larger statistical uncertainty, but begins to show signs of deviation from the original KK-matrix in the energy range outside of the fit region. Parameterization 𝖢\mathsf{C} does not have sufficient flexibility to describe the complete energy dependence - while it does correctly reproduce the resonance shape in δϕ\delta^{\phi} and the presence of a dip in η\eta, the precise energy dependence of η\eta is not correct and it fails to describe δσ\delta^{\sigma} at high energies, away from the energy region where the pole dominates. Parameterisation 𝖣\mathsf{D} introduces too much parameter-space freedom for the limited set of data points available. As such we see rapid energy variation that is not really required and a large degree of statistical uncertainty. Parameterisation 𝖤\mathsf{E} shows that a precise knowledge of the form of the “background” is not required to reproduce the energy dependence in a limited region - a second KK-matrix pole (at higher energy) is able to mock-up the polynomial behavior away from the resonance pole quite well.

Our principal interest lies in identifying resonances as poles in the complex-ss plane – analytically continuing the fitted model amplitudes we find that all five have single poles on sheets 𝖨𝖨\mathsf{II} and 𝖨𝖨𝖨\mathsf{III} whose locations are in a rather good agreement with the input pole position (see Fig. 5). The residues at the pole agree similarly. The statistical precision of the pole position determination typically decreases as we introduce more parameters into the model. The pleasing observation here is that in order to determine the position of a sharp resonance we do not need to have precise knowledge of the form of the energy-dependence of the “background”.

In summary, the “pointwise” strategy may provide a less model-dependent approach for extracting phase shifts and inelasticities, however, this method is limited by the number of points for which accidental degeneracies appear. Parameterizing the scattering amplitude allow us to make use of all measured energy levels, however we need to find suitable parameterizations. One strategy is to explore the “pointwise” approach to find a crude guide to the energy dependence and then build parameterizations which are able to reproduce the obtained form. The parameterizations, which should respect certain constraints applicable to scattering amplitudes, can be made progressively more sophisticated in an effort to reduce the overall χ2\chi^{2} – in this sense the approach is not dissimilar to what is done with real experimental data.

V Summary

Using the Hamlitonian formalism applied to a model of interacting relativistic fields, we derived a generalized Lüscher’s formula Lusher:1991 ; Gottlieb:1995 ; Liu:2005 ; Doring:2011 for two-particle scattering, in both the single- and coupled-channel systems, in moving frames.

Our results are consistent with the ones obtained previously in Lusher:1991 ; Gottlieb:1995 ; Liu:2005 ; Doring:2011 . In the coupled-channel case we are challenged by the fact that, even for dominance of a single partial-wave, the system is underconstrained for determination of multiple-channel phase-shifts and inelasticities from a single determined finite-volume energy level. Using a toy model of two-channel SS-wave scattering we demonstrated that it is possible to determine this information if multiple energy levels are determined.

Two possible strategies for extracting information from discrete spectra of a coupled-channel system were discussed, one approach utilizes the near degeneracy of energy levels in different volumes and total momenta of system, and another fits the discrete spectra by parameterizating the scattering amplitudes with certain numbers of parameters. These strategies may be useful for the analysis of future lattice QCD data. In particular, the coupled-channel analysis has to be considered for the strongly coupled systems, for instance, ππ,KK¯\pi\pi,K\bar{K} and ηη\eta\eta system in SS-wave.

VI ACKNOWLEDGMENTS

We thank D. B. Renner and M. R. Pennington for useful discussions, and our colleagues within the Hadron Spectrum Collaboration for their continued assistance. PG, RGE and JJD acknowledge support from U.S. Department of Energy contract DE-AC05-06OR23177, under which Jefferson Science Associates, LLC, manages and operates Jefferson Laboratory. JJD also acknowledges the support of the U.S. Department of Energy Early Career award contract DE-SC0006765. APS acknowledges the support of the U.S. Department of Energy grant under contract DE-FG0287ER40365.

Appendix A Relativistic Lippmann-Schwinger equation from Hamiltonian formalism

Following the method presented in Glazek:1993 ; Bakker:2003 , we treat the relativistic dynamics of particle scattering in the Hamiltonian formalism approach. We start from the covariant Lagrangian in Eq.(1) and choose to quantize the field operators in the instant form Dirac:1949 ; the construction of generators of the Poincaré group can be done in a standard way in quantum field theory. In principle, one needs to solve eigenstate equations H^|Ψ=E|Ψ\hat{H}|\Psi\rangle=E|\Psi\rangle on a instant quantization plane, where |Ψ|\Psi\rangle denotes the Poincaré covariant state vector spanning the complete Fock space. We truncate the Fock space up to three-body states, assuming that this is sufficient to describe low-energy physics. Thus, the eigenstate equations reduce to a matrix equation given in Eq.(2). Eliminating the three-body sector, we end up with the relativistic Schrödinger-like equation for a two-body system given in Eq.(3). For simplicity, we have assumed that the two charged scalars scattering have equal mass, however, the conclusion of this work can be generalized to non-equal masses case as well (c.f. Fu:2012 ; Prelovsek:2012 ).

We choose the center-of-mass frame of the many-body system to construct multiple-particle states |JM|JM\rangle having total spin JJ and spin projection MM. The two-particle state |ϕ+ϕ|\phi^{+}\phi^{-}\rangle in the center-of-mass frame is given by

|ϕ+ϕ;JM\displaystyle\big{|}\phi^{+}\phi^{-};JM\big{\rangle} =2sd3𝐩1(2π)32Ep1d3𝐩2(2π)32Ep2δ3(𝐩1+𝐩2)\displaystyle=2\sqrt{s}\int\!\!\frac{d^{3}\mathbf{p}_{1}}{(2\pi)^{3}2E_{p_{1}}}\frac{d^{3}\mathbf{p}_{2}}{(2\pi)^{3}2E_{p_{2}}}\delta^{3}(\mathbf{p}_{1}+\mathbf{p}_{2})
×(2π)3φJM(2)(𝐩1,𝐩2)a𝐩1b𝐩2|0,\displaystyle\quad\quad\quad\quad\times(2\pi)^{3}\varphi^{(2)}_{JM}(\mathbf{p}_{1},\mathbf{p}_{2})\,a^{{\dagger}}_{\mathbf{p}_{1}}b^{{\dagger}}_{\mathbf{p}_{2}}\,\big{|}0\big{\rangle},

where 𝐩i\mathbf{p}_{i} is the momentum of the ii-th particle and s\sqrt{s} is the invariant mass of the two-particle system. φJM(2)(𝐩1,𝐩2)\varphi^{(2)}_{JM}(\mathbf{p}_{1},\mathbf{p}_{2}) is the wavefunction of the two-particle system, describing the momentum distribution of the two particles.

Similarly, the three-particle state |ϕ+ϕθ|\phi^{+}\phi^{-}\theta\rangle is given by

|ϕ+ϕθ;JM=2s\displaystyle\big{|}\phi^{+}\phi^{-}\theta;JM\big{\rangle}=2\sqrt{s} d3𝐩1(2π)32Ep1d3𝐩2(2π)32Ep2d3𝐩3(2π)32Ep3\displaystyle\int\!\!\!\frac{d^{3}\mathbf{p}_{1}}{(2\pi)^{3}2E_{p_{1}}}\frac{d^{3}\mathbf{p}_{2}}{(2\pi)^{3}2E_{p_{2}}}\frac{d^{3}\mathbf{p}_{3}}{(2\pi)^{3}2E_{p_{3}}}
×(2π)3δ3(𝐩1+𝐩2+𝐩3)\displaystyle\quad\times(2\pi)^{3}\delta^{3}(\mathbf{p}_{1}+\mathbf{p}_{2}+\mathbf{p}_{3})
×φJM(3)(𝐩1,𝐩2,𝐩3)a𝐩1b𝐩2d𝐩3|0,\displaystyle\quad\times\varphi^{(3)}_{JM}(\mathbf{p}_{1},\mathbf{p}_{2},\mathbf{p}_{3})\,a^{{\dagger}}_{\mathbf{p}_{1}}b^{{\dagger}}_{\mathbf{p}_{2}}d^{{\dagger}}_{\mathbf{p}_{3}}\big{|}0\big{\rangle},

where φJM(3)(𝐩1,𝐩2,𝐩3)\varphi^{(3)}_{JM}(\mathbf{p}_{1},\mathbf{p}_{2},\mathbf{p}_{3}) is the wavefunction of the three-particle system. The wavefunctions are normalized so that the normalization of states is JM|JM=2s(2π)3δ3(𝟎)\langle JM|JM\rangle=2\sqrt{s}\,(2\pi)^{3}\,\delta^{3}(\mathbf{0}).

It is straightforward to evaluate the matrix elements of the eigenstate equations Eq.(2) and to get coupled equations for the wavefunctions

φJM(2)(𝐪)=gs2𝐪2+m2d3𝐤(2π)32𝐤2+μ2[φJM(3)(𝐪+12𝐤,𝐤)2(𝐪+𝐤)2+m2+φJM(3)(𝐪12𝐤,𝐤)2(𝐪𝐤)2+m2],\displaystyle\varphi^{(2)}_{JM}(\mathbf{q})=\frac{g}{\sqrt{s}-2\sqrt{\mathbf{q}^{2}+m^{2}}}\int\!\!\!\frac{d^{3}\mathbf{k}^{\prime}}{(2\pi)^{3}2\sqrt{\mathbf{k^{\prime}}^{2}+\mu^{2}}}\left[\frac{\varphi^{(3)}_{JM}(\mathbf{q}+\tfrac{1}{2}\mathbf{k^{\prime}},\mathbf{k^{\prime}})}{2\sqrt{(\mathbf{q}+\mathbf{k^{\prime}})^{2}+m^{2}}}+\frac{\varphi^{(3)}_{JM}(\mathbf{q}-\tfrac{1}{2}\mathbf{k^{\prime}},\mathbf{k^{\prime}})}{2\sqrt{(\mathbf{q}-\mathbf{k^{\prime}})^{2}+m^{2}}}\right],
φJM(3)(𝐪,𝐤)=g12(12𝐤𝐪)2+m2φJM(2)(𝐪12𝐤)+12(12𝐤+𝐪)2+m2φJM(2)(𝐪+12𝐤)s(12𝐤+𝐪)2+m2(12𝐤𝐪)2+m2𝐤2+μ2,\displaystyle\varphi^{(3)}_{JM}(\mathbf{q},\mathbf{k})=g\;\frac{\frac{1}{2\sqrt{(\frac{1}{2}\mathbf{k}-\mathbf{q})^{2}+m^{2}}}\,\varphi^{(2)}_{JM}(\mathbf{q}-\frac{1}{2}\mathbf{k})+\frac{1}{2\sqrt{(\frac{1}{2}\mathbf{k}+\mathbf{q})^{2}+m^{2}}}\,\varphi^{(2)}_{JM}(\mathbf{q}+\tfrac{1}{2}\mathbf{k})}{\sqrt{s}-\sqrt{(\frac{1}{2}\mathbf{k}+\mathbf{q})^{2}+m^{2}}-\sqrt{(\tfrac{1}{2}\mathbf{k}-\mathbf{q})^{2}+m^{2}}-\sqrt{\mathbf{k}^{2}+\mu^{2}}},

where we have used a short-hand notation for wavefunctions φJM(2)(𝐪)\varphi^{(2)}_{JM}(\mathbf{q}) and φJM(3)(𝐪,𝐤)\varphi^{(3)}_{JM}(\mathbf{q},\mathbf{k}), with arguments of relative momenta defined by 𝐪=12(𝐩1𝐩2),𝐤=𝐩3\mathbf{q}=\frac{1}{2}(\mathbf{p}_{1}-\mathbf{p}_{2}),\mathbf{k}=-\mathbf{p}_{3}. Eliminating the three-body wavefunction, we get a relativistic equation for the two-particle state |ϕ+ϕ|\phi^{+}\phi^{-}\rangle with an effective non-local potential generated from the neutral scalar exchange between two charged scalars

φJM(2)(𝐪)=1s2𝐪2+m2d3𝐤(2π)3V(𝐪,𝐤)φJM(2)(𝐤),\displaystyle\varphi_{JM}^{(2)}(\mathbf{q})=\frac{1}{\sqrt{s}-2\sqrt{\mathbf{q}^{2}+m^{2}}}\int\!\!\frac{d^{3}\mathbf{k}}{(2\pi)^{3}}\,V(\mathbf{q},\mathbf{k})\,\varphi_{JM}^{(2)}(\mathbf{k}),

with

V(𝐪,𝐤)=g2411Σ(𝐪)1(𝐤2+m2)1(𝐤𝐪)2+μ21s𝐤2+m2𝐪2+m2(𝐤𝐪)2+μ2,V(\mathbf{q},\mathbf{k})=\frac{g^{2}}{4}\frac{1}{1-\Sigma(\mathbf{q})}\frac{1}{(\mathbf{k}^{2}+m^{2})}\frac{1}{\sqrt{(\mathbf{k}-\mathbf{q})^{2}+\mu^{2}}}\frac{1}{\sqrt{s}-\sqrt{\mathbf{k}^{2}+m^{2}}-\sqrt{\mathbf{q}^{2}+m^{2}}-\sqrt{(\mathbf{k}-\mathbf{q})^{2}+\mu^{2}}}, (26)

where the self-energy contribution is,

Σ(𝐪)=1s2𝐪2+m2g24d3𝐤(2π)3\displaystyle\Sigma(\mathbf{q})=\frac{1}{\sqrt{s}-2\sqrt{\mathbf{q}^{2}+m^{2}}}\frac{g^{2}}{4}\;\int\!\!\!\frac{d^{3}\mathbf{k}^{\prime}}{(2\pi)^{3}} 1s𝐤2+m2𝐪2+m2(𝐤𝐪)2+μ2\displaystyle\frac{1}{\sqrt{s}-\sqrt{\mathbf{k^{\prime}}^{2}+m^{2}}-\sqrt{\mathbf{q}^{2}+m^{2}}-\sqrt{(\mathbf{k^{\prime}}-\mathbf{q})^{2}+\mu^{2}}}
×1𝐪2+m21𝐤2+m21(𝐤𝐪)2+μ2.\displaystyle\times\frac{1}{\sqrt{\mathbf{q}^{2}+m^{2}}}\frac{1}{\sqrt{\mathbf{k^{\prime}}^{2}+m^{2}}}\frac{1}{\sqrt{(\mathbf{k^{\prime}}-\mathbf{q})^{2}+\mu^{2}}}.

In coordinate space, the wave-equation becomes

ψJM(𝐫)=d3𝐫G0(𝐫𝐫,s)d3𝐳V~(𝐫,𝐳)ψJM(𝐳),\displaystyle\psi_{JM}(\mathbf{r})=\int\!\!d^{3}\mathbf{r}^{\prime}\,G_{0}(\mathbf{r}-\mathbf{r^{\prime}},\sqrt{s})\int\!\!d^{3}\mathbf{z}\,\widetilde{V}(\mathbf{r^{\prime}},-\mathbf{z})\,\psi_{JM}(\mathbf{z}),

where ψJM(𝐫)\psi_{JM}(\mathbf{r}) and V~(𝐫,𝐳)\widetilde{V}(\mathbf{r^{\prime}},-\mathbf{z}) are the Fourier transforms of the momentum-space two-particle wavefunction and effective potential, respectively. The free Green’s function is given in Eq.(II.1). Performing the angular integral, the free Green’s function reads

G0(𝐫,s)=12irqdq(2π)2eiqreiqrs2q2+m2,\displaystyle G_{0}(\mathbf{r},\sqrt{s})=\frac{1}{2ir}\int\limits_{-\infty}^{\infty}\!\!\frac{qdq}{(2\pi)^{2}}\frac{e^{iqr}-e^{-iqr}}{\sqrt{s}-2\sqrt{q^{2}+m^{2}}}, (27)

which has the following singularities in the complex qq plane: two poles on the real axis, q=±kq=\pm k, and two branch cuts on the imaginary axis ±[im,i]\pm\left[im,i\infty\right], see Fig.6. We choose the contour C1+C2C_{1}+C_{2} to include the pole at q=kq=k and circle around the cut [im,i]\left[im,i\infty\right] on the upper half-plane for first term with factor eikre^{ikr} and choose the contour C1+C3C_{1}+C_{3} to include the pole at q=kq=-k and circle around the cut [im,i]\left[-im,-i\infty\right] on the lower half-plane for the second term with factor eikre^{-ikr}, as shown in Fig.6. The contour integral leads to

G0(𝐫,s)\displaystyle G_{0}(\mathbf{r},\sqrt{s}) =s2eikr4πr1rmρdρ(2π)2ρ2m2eρrk2+ρ2,\displaystyle=-\frac{\sqrt{s}}{2}\frac{e^{ikr}}{4\pi r}-\frac{1}{r}\int\limits_{m}^{\infty}\!\!\frac{\rho d\rho}{(2\pi)^{2}}\sqrt{\rho^{2}-m^{2}}\frac{e^{-\rho r}}{k^{2}+\rho^{2}},

where k=12s4m2k=\tfrac{1}{2}\sqrt{s-4m^{2}} is the momentum of either particle in the rest frame of the two-particle system. The first term on the right hand side comes from the poles at q=±kq=\pm k and is proportional to the usual non-relativistic Green’s function which oscillates over the path of propagation. The second term comes from the contribution of the discontinuity crossing the branch cuts at ±[im,i]\pm\left[im,i\infty\right] - it decays exponentially over the propagation. Expanding ρ2m2=ρ(1𝒪(m2ρ2))\sqrt{\rho^{2}-m^{2}}=\rho\left(1-\mathcal{O}\big{(}\frac{m^{2}}{\rho^{2}}\big{)}\right), at large separations, the free Green’s function may be approximated by

G0(𝐫;s)s214πr[eikr+2πemrrs],\displaystyle G_{0}(\mathbf{r};\sqrt{s})\approx-\frac{\sqrt{s}}{2}\,\frac{1}{4\pi r}\left[e^{ikr}+\frac{2}{\pi}\frac{e^{-mr}}{r\sqrt{s}}\right], (28)

and therefore the exponential decaying term can be dropped in the limit rm1r\gg m^{-1}.

Refer to caption
Figure 6: Integration contours and singularities of the free Green’s function in Eq.(27) on the complex qq plane.

Appendix B Expansion of Green’s function and regularization of expansion coefficients

We start from the expansion of the Green’s function

1L3\displaystyle\frac{1}{L^{3}} 𝐪𝐏𝐐ei𝐪𝐫k2𝐪2\displaystyle\sum_{\mathbf{q}\in\mathbf{P}_{\mathbf{Q}}}\frac{e^{i\mathbf{q}\cdot\mathbf{r}}}{k^{2}-\mathbf{q}^{2}}
=k4πn0(kr)jmjgjmj(𝐐)(k)jj(kr)Yjmj(𝐫^),\displaystyle=\frac{k}{4\pi}\,n_{0}(kr)-\sum_{jm_{j}}g^{(\mathbf{Q})}_{jm_{j}}\!(k)\,j_{j}(kr)\,Y_{jm_{j}}(\mathbf{\hat{r}}),

where the summation of 𝐪\mathbf{q} runs over 𝐏𝐐={𝐪3|𝐪=2πL𝐧+𝐐, for 𝐧3}\mathbf{P}_{\mathbf{Q}}=\{\mathbf{q}\in\mathbb{R}^{3}|\mathbf{q}=\frac{2\pi}{L}\mathbf{n}+\mathbf{Q},\mbox{ for }\mathbf{n}\in\mathbb{Z}^{3}\}. The expansion coefficients are given by Lusher:1991

gjmj(𝐐)(k)=4πL3𝐪𝐏𝐐ijqjkjYjmj(𝐪^)𝐪2k2δj0δmj04π1r|r0.\displaystyle g^{(\mathbf{Q})}_{jm_{j}}\!(k)=\frac{4\pi}{L^{3}}\sum_{\mathbf{q}\in\mathbf{P}_{\mathbf{Q}}}\!i^{j}\,\frac{q^{j}}{k^{j}}\,\frac{Y_{jm_{j}}^{*}\!(\mathbf{\hat{q}})}{\mathbf{q}^{2}-k^{2}}-\frac{\delta_{j0}\,\delta_{m_{j}0}}{\sqrt{4\pi}}\left.\frac{1}{r}\right|_{r\rightarrow 0}.

Note that the definition of nj(x)n_{j}(x) in this work differs from the definition in Lusher:1991 by a overall negative sign.

Using the identities

jj\displaystyle j_{j} (k|𝐫𝐫|)Yjmj(𝐫𝐫^)\displaystyle\!\big{(}k|\mathbf{r}-\mathbf{r^{\prime}}|\big{)}\,Y_{jm_{j}}(\widehat{\mathbf{r}-\mathbf{r^{\prime}}})
=r<r4πlmllmlilljjl(kr)jl(kr)Ylml(𝐫^)Ylml(𝐫^)\displaystyle\stackrel{{\scriptstyle r^{\prime}<r}}{{=}}\sqrt{4\pi}\sum_{\begin{subarray}{c}lm_{l}\\ l^{\prime}m_{l^{\prime}}\end{subarray}}i^{l-l^{\prime}-j}\,j_{l}(kr)\,j_{l^{\prime}}(kr^{\prime})\,Y_{lm_{l}}(\mathbf{\hat{r}})\,Y^{*}_{l^{\prime}m_{l^{\prime}}}(\mathbf{\hat{r}^{\prime}})
×(2j+1)(2l+1)2l+1jmj;lml|lmlj0;l0|l0,\displaystyle\hskip 56.9055pt\times\sqrt{\tfrac{(2j+1)(2l^{\prime}+1)}{2l+1}}\,\langle jm_{j};l^{\prime}m_{l^{\prime}}|lm_{l}\rangle\,\langle j0;l^{\prime}0|l0\rangle,

and

k4πn0(k|𝐫𝐫|)=r<rklmlnl(kr)jl(kr)Ylml(𝐫^)Ylml(𝐫^),\displaystyle\frac{k}{4\pi}\,n_{0}\!\big{(}k|\mathbf{r}-\mathbf{r^{\prime}}|\big{)}\stackrel{{\scriptstyle r^{\prime}<r}}{{=}}k\sum_{lm_{l}}n_{l}(kr)\,j_{l}(kr^{\prime})\,Y_{lm_{l}}(\mathbf{\hat{r}})\,Y^{*}_{lm_{l}}(\mathbf{\hat{r}}^{\prime}),

we also have

1L3\displaystyle\frac{1}{L^{3}} 𝐪𝐏𝐐ei𝐪(𝐫𝐫)k2𝐪2\displaystyle\!\sum_{\mathbf{q}\in\mathbf{P}_{\mathbf{Q}}}\frac{e^{i\mathbf{q}\cdot(\mathbf{r}-\mathbf{r^{\prime}})}}{k^{2}-\mathbf{q}^{2}}
=r<rlmllml[δlml,lmlnl(kr)lml,lml(𝐐)(k)jl(kr)],\displaystyle\stackrel{{\scriptstyle r^{\prime}<r}}{{=}}\sum_{\begin{subarray}{c}lm_{l}\\ l^{\prime}m_{l^{\prime}}\end{subarray}}\left[\delta_{l^{\prime}m^{\prime}_{l^{\prime}},lm_{l}}\,n_{l}(kr)-\mathcal{M}^{(\mathbf{Q})}_{l^{\prime}m^{\prime}_{l^{\prime}},lm_{l}}(k)\,j_{l}(kr)\right],
×kjl(kr)Ylml(𝐫^)Ylml(𝐫^),\displaystyle\quad\quad\quad\quad\quad\times k\,j_{l^{\prime}}(kr^{\prime})\,Y_{lm_{l}}(\mathbf{\hat{r}})\,Y^{*}_{l^{\prime}m_{l^{\prime}}}(\mathbf{\hat{r}^{\prime}}), (30)

with

\displaystyle\mathcal{M} (k)lml,lml(𝐐)=jmjillj4πkgjmj(𝐐)(k){}^{(\mathbf{Q})}_{l^{\prime}m^{\prime}_{l^{\prime}},lm_{l}}(k)=\sum_{jm_{j}}i^{l-l^{\prime}-j}\frac{\sqrt{4\pi}}{k}g^{(\mathbf{Q})}_{jm_{j}}(k)
×(2j+1)(2l+1)2l+1jmj;lml|lmlj0;l0|l0.\displaystyle\times\sqrt{\tfrac{(2j+1)(2l^{\prime}+1)}{2l+1}}\langle jm_{j};l^{\prime}m_{l^{\prime}}|lm_{l}\rangle\langle j0;l^{\prime}0|l0\rangle. (31)

If 𝐐\bf{Q} is identified with 12γ𝐏\tfrac{1}{2\gamma}\bf{P} for two-equal-mass-particle scattering, the generalization to two-unequal-mass-particle scattering leads to 12γ𝐏(1+m12m22E2)\frac{1}{2\gamma}\mathbf{P}(1+\frac{m_{1}^{2}-m_{2}^{2}}{E^{2}}) Fu:2012 ; Prelovsek:2012 , these expressions are the same as those presented in Gottlieb:1995 , and the regularisation procedure outlined therein can be followed. Once the Lorentz contraction is considered (substitution of box volume L3L^{3} and momentum 2πL𝐧,𝐧3\frac{2\pi}{L}\mathbf{n},\mathbf{n}\in\mathbb{Z}^{3} by γL3\gamma L^{3} and 2πLγ1𝐧\frac{2\pi}{L}\gamma^{-1}\mathbf{n} respectively), the function gjmj(𝐐)(k)g^{(\mathbf{Q})}_{jm_{j}}\!(k) is related to the Zeta function defined in Eq.(93) of Gottlieb:1995 by

gjmj(𝐐)(k)=1π1γLij(kL2π)jZjmj𝐝(1,kL2π).\displaystyle g^{(\mathbf{Q})}_{jm_{j}}\!(k)=\frac{1}{\pi}\frac{1}{\gamma L}\frac{i^{j}}{\left(\frac{kL}{2\pi}\right)^{j}}Z^{\mathbf{d}*}_{jm_{j}}\left(1,\tfrac{kL}{2\pi}\right). (32)

References

  • (1) C. Michael, Nucl. Phys. B 259, 58 (1985).
  • (2) M. Luscher and U. Wolff, Nucl. Phys. B 339, 222 (1990).
  • (3) B. Blossier, M. Della Morte, G. von Hippel, T. Mendes and R. Sommer, JHEP 0904, 094 (2009)
  • (4) J. J. Dudek et al. (Hadron Spectrum Collaboration), Phys. Rev. D 82, 034508 (2010).
  • (5) R. G. Edwards, J. J. Dudek, D. G. Richards, and S. J. Wallace, Phys. Rev. D 84, 074508 (2011).
  • (6) J. J. Dudek et al. (Hadron Spectrum Collaboration), Phys. Rev. D 86, 034031 (2012).
  • (7) L. Maiani and M. Testa, Phys. Lett. B 245, 585 (1990).
  • (8) M. Lüscher, Nucl. Phys. B 354, 531 (1991).
  • (9) K. Rummukainen, S. Gottlieb, Nucl. Phys. B 450, 397 (1995).
  • (10) C.-J.D. Lin, G. Martinelli, C. T. Sachrajda and M. Testa, Nucl. Phys. B 619, 467 (2001).
  • (11) N. H. Christ, C. Kim and T.Yamazaki, Phys. Rev. D 72, 114506 (2005).
  • (12) V. Bernard, Ulf-G. Meißner and A. Rusetsky, Nucl. Phys. B 788, 1 (2008).
  • (13) V. Bernard, M. Lage, Ulf-G. Meißner and A.Rusetsky, JHEP 0808, 024 (2008).
  • (14) S. He, X. Feng, C. Liu, JHEP 0507, 011 (2005).
  • (15) M. Döring, Ulf-G. Meißner,E. Oset and A. Rusetsky, Eur. Phys. J. A 47, 139 (2011)
  • (16) S. Aoki et al. [HAL QCD Collaboration], Proc. Japan Acad. B 87, 509 (2011)
  • (17) R. A. Briceno and Z. Davoudi, arXiv:1204.1110 [hep-lat].
  • (18) M. T. Hansen and S. R. Sharpe, Phys. Rev. D 86, 016007 (2012)
  • (19) S. Aoki et al. [CP-PACS Collaboration], Phys. Rev. D 76, 094506 (2007)
  • (20) K. Sasaki, and N. Ishizuka, Phys. Rev. D 78, 014511 (2008).
  • (21) X. Feng, K. Jansen, and D. B. Renner, Phys. Rev. D 83, 094505 (2011).
  • (22) J. J. Dudek et al. (Hadron Spectrum Collaboration), Phys. Rev. D 83, 071504 (2011).
  • (23) S. R. Beane et al. (NPLQCD Collaboration), Phys. Rev. D 85, 034505 (2012).
  • (24) C. B. Lang, D. Mohler, S. Prelovsek and M. Vidmar, Phys. Rev. D 84, 054503 (2011).
  • (25) S. Aoki et al. [CS Collaboration], Phys. Rev. D 84, 094505 (2011)
  • (26) S. Glazek, A. Harindranath, S. Pinsky, J. Shigemitsu, and K. Wilson, Phys. Rev. D 47, 1599 (1993).
  • (27) B.L.G. Bakker, M. van Iersel, F. Pijlman, Few Body Syst. 33, 27 (2003).
  • (28) P. A. M. Dirac, Rev. Mod. Phys. 21, 392 (1949).
  • (29) C. E. Thomas, R. G. Edwards and J. J. Dudek, Phys. Rev. D 85, 014507 (2012).
  • (30) A. M. Torres, M. Bayar, D. Jido and E. Oset, Phys. Rev. C 86, 055201 (2012).
  • (31) J. L. Basdevant and E. L. Berger, Phys. Rev. D 16, 657 (1977).
  • (32) J. J. Dudek, R. G. Edwards and C. E. Thomas, arXiv:1212.0830 [hep-ph].
  • (33) Z. Fu , Phys. Rev. D 85, 014506 (2012).
  • (34) L. Leskovec and S. Prelovsek, Phys. Rev. D 85, 114507 (2012).