Discontinuity calculus and applications to two-body coupled-channel scattering

Hao-Jie Jing , Xiong-Hui Cao , Feng-Kun Guo

Front. Phys. ›› 2026, Vol. 21 ›› Issue (5) : 056201

PDF (1417KB)
Front. Phys. ›› 2026, Vol. 21 ›› Issue (5) : 056201 DOI: 10.15302/frontphys.2026.056201
RESEARCH ARTICLE

Discontinuity calculus and applications to two-body coupled-channel scattering

Author information +
History +
PDF (1417KB)

Abstract

We present a novel method, termed discontinuity calculus, for computing discontinuities of complex functions. This framework enables a systematic investigation of both analytic continuation and the topological structure of Riemann surfaces. We apply this calculus to analyze the analytic continuation of partial-wave amplitudes in two-body coupled-channel scattering problems and discuss their uniformization of the corresponding Riemann surfaces. This methodology offers new perspectives and tools for analyzing coupled-channel scattering problems in quantum scattering theory.

Graphical abstract

Keywords

discontinuity calculus / quantum scattering theory

Cite this article

Download citation ▾
Hao-Jie Jing, Xiong-Hui Cao, Feng-Kun Guo. Discontinuity calculus and applications to two-body coupled-channel scattering. Front. Phys., 2026, 21(5): 056201 DOI:10.15302/frontphys.2026.056201

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

The analytic S-matrix theory is one of the most powerful tools for studying scattering problems. S-matrix elements, or scattering amplitudes, are functions of the invariant masses involved in reaction processes. These elements provide the invariant mass spectral distributions that can be directly compared with experimental data. In quantum field theory, bound or resonance states correspond to poles in scattering amplitudes. In hadron physics, particularly, the locations and residues of the poles contain crucial information about hadron properties; e.g., the branching fractions of decays of a resonance can be unambiguously defined through the residues of the poles [1]. Causality constrains the unstable resonance poles to be on unphysical Riemann sheets of the complex energy plane, preventing their appearance on the physical sheet. Thus, it is necessary to perform analytic continuation of S-matrix elements, extending their domains from the real axis to the complex plane, to locate such poles.

In this work, we propose a novel method, termed discontinuity calculus, for analytic continuation of complex functions and visualization of their Riemann surface structures. This approach provides a systematic way to derive the analytic continuation of any complex function along its branch cuts, as well as the structure of the corresponding Riemann surface. By applying this methodology, we have analyzed the analytic continuation and Riemann surface structure of the partial-wave scattering matrix in two-body coupled-channel problems [2, 3]. Furthermore, we establish the connection between the uniformization problem of the partial-wave scattering matrix and the uniformization theorem in complex analysis.

This paper is organized as follows. In Section 2, we introduce the discontinuity calculus. In Section 3, we employ this framework to systematically investigate the analytic continuation of the scalar two-point Green’s function and the topological structure of the associated Riemann surface. In Section 4, we extend this analysis to the analytic continuation and Riemann surface topology of the partial-wave scattering matrix in two-body coupled-channel problems. In Section 5, we address the uniformization problem of the partial-wave scattering matrix. Finally, Section 6 provides a summary.

2 Discontinuity calculus

In this section, we develop a formalism for computing discontinuities of complex functions.

2.1 Definitions

For any complex function f(z), F(ω), and constant α, we define the discontinuity operator D, which complies with the following properties):

(i) Vanishing discontinuity of a holomorphic function h(z)

Dh(z)=0.

(ii) Linearity

D[α1f1(z)+α2f2(z)]=α1Df1(z)+α2Df2(z).

(iii) Leibniz rule for the discontinuity of a product of two functions

D[f1(z)f2(z)]=Df1(z)f2(z)+f1(z)Df2(z)Df1(z)Df2(z).

(iv) Chain rule for the discontinuity of a composite function F[f(z)]

DF[f(z)]=F[f(z)][F(ω)DF(ω)]ω=f(z)Df(z).

In particular, if DF(ω)=Df(z)=0, then DF[f(z)]=0; if DF(ω)=0 and Df(z)0, then DF[f(z)]=F[f(z)]F[f(z)Df(z)]; if DF(ω)0 and Df(z)=0, then DF[f(z)]=DF(ω)|ω=f(z).

Using (R1, R3), one can prove that the discontinuities of the functions f(z) and 1/f(z) satisfy the following relation:

Df(z)=f(z)D[1/f(z)][f(z)Df(z)].

In general, the discontinuity of a complex function f(z) can be decomposed into the following form:

Df(z)=i=1nKif(z)θi(z)+j=1nαjδj(z).

The right-hand side of Eq. (2) comprises two types of contributions: branch cuts and poles. The first term describes the branch cuts of f(z), which are generally expressed using the Heaviside θ function. Here, they are represented by θi(z)(i=1,,n) in simplified notation, where n denotes the number of cuts. The specific form of the argument within the θ function depends on the particular cut. The components Kif(z) determine the analytic continuation of f(z). In the second term, the components δj(z)(j=1,,n) correspond to the poles of f(z), where n is the number of poles. The discontinuity arising from poles is typically expressed using the Dirac δ function and its derivatives; the coefficients αj are determined by the Laurent series expansion of f(z). We refer to Kif(z) as the continuation kernel of f(z), and Ki as the continuation kernel operator. For a given cut of a complex function f(z) or a composite function F[f(z)], represented by θi(z), it can be readily shown that the corresponding continuation kernel operator Ki also satisfies (R1)–(R4).

Furthermore, we can define the continuation generator Ti1Ki. The action of the continuation generator Ti on the function f(z), denoted Tif(z), yields the expression of f(z) after analytic continuation across the cut specified by θi(z). We can then calculate the discontinuity of the function Tif(z):

D[Tif(z)]=j=1nKj[Tif(z)]θj(z)+j=1nαjδj(z),

which leads to the expression of the function Tif(z) after analytic continuation across the cut of Tif(z) specified by θj(z), namely TjTif(z). By repeating this process, we can obtain the expression of f(z) on all Riemann sheets after analytic continuation, and thus determine the structure of the Riemann surface RS{f(z)}.

2.2 Examples

Let us now examine some instructive examples.

Example 1: the square root function f(z)=z (zC). Using (R3) to expand Df2(z)=0, the following equation is obtained:

(Dz2z)Dz=0.

The square root function has two branch points at 0 and , and any continuous curve connecting these branch points can serve as a branch cut. When the branch cut is chosen along the positive real axis, the following equation holds:

Dz=2z×θ(z)+0×θ(z)=2zθ(z),

where the two terms following the first equality correspond to the two solutions of Eq. (4). Thus, the continuation kernel for the square root function is K1z=2z, with the trivial case K2z=0 omitted.

The continuation generator is given by

T1z=(1K1)z=z.

Since T1z and z differ only by a sign, it follows that D(T1z)=Dz. Therefore, the analytic continuation of T1z does not introduce new continuation generators. Furthermore, since T12z=T1z=z, this indicates that the order of T1 is 2, generating a cyclic group of order 2, denoted Z2.

In summary, the Riemann surface of the square root function RS{z} consists of two Riemann sheets:

RS{z}=C×{T1T121}C×Z2.

Moreover, it can also be easily shown that RS{zn}C×Zn for nN.

Example 2: the logarithmic function f(z)=logz(zCC{0}). Using (R4) to expand Delogz=0, the following equation is obtained:

eDlogz=1.

The logarithmic function also has two branch points located at 0 and . If the branch cut is chosen along the negative real axis, the above equality leads to the discontinuity of the logarithm:

Dlogz=2mπiθ(z),mZ.

From this expression, the continuation kernel is given by Kmlogz=2mπi.

The continuation generator is given by

Tmlogz=(1Km)logz=logz2mπi.

Since D(Kmlogz)=0, no new continuation generators are needed for the analytic continuation of Tmlogz. Furthermore, group multiplication is expressed as Tm1(Tm2logz)=Tm1+m2logz. This result establishes that the continuation generators T±1 alone suffice to construct the complete analytic continuation. Consequently, the discontinuity of the logarithmic function can be rewritten as

Dlogz=±2πiθ(z).

The sign +() on the right-hand side corresponds to the result obtained by analytic continuation of the logarithmic function along the upper (lower) edge of the branch cut.

In summary, the Riemann surface of the logarithmic function, RS{logz}, consists of countably infinite Riemann sheets and can be expressed as

RS{logz}=C×{T1,T1T1T11}C×Z.

Example 3: the meromorphic function f(z)=1/z (zC). By taking the derivative of both sides of Eq. (6) with respect to z and interchanging the order of the derivative operator and the discontinuity operator, which is valid due to the linearity property (R2), one obtains

D(1z)=2πiδ(z),

where the sign (+) corresponds to the variable z approaching zero along the upper (lower) edge of the real axis, namely Imz>0 (Imz<0). Equation (7) indicates that the function 1/z has only a single pole at z=0. Furthermore, we have the following identities (nN):

D(1zn+1)=2πi(1)nn!dndznδ(z).

3 Riemann surface of two-point Green’s function

In this section, we employ the discontinuity calculus formalism developed in Section 2 to systematically investigate the Riemann surface topology of the scalar two-point one-loop Green’s function,

G(p2)R4id4q(2π)41(q+2m12+iϵ)(q2m22+iϵ),

where q±=p/2±q, p is the four-momentum of the external line, m1 and m2 are the masses of the intermediate particles, and ϵ=0+ is a positive infinitesimal, which is used to determine the way of integration path bypassing the poles of the integrand.

First, we demonstrate that the discontinuity of the Green’s function G(s)(wheres=p2) can be explicitly derived solely from its definition in Eq. (8). By applying the discontinuity calculus in Eq. (7), the discontinuity of G(s) in the physical region, which is defined as the upper edge of the real axis on the first Riemann sheet of the complex s plane, can be obtained from the following substitution in Eq. (8)):

1(q+2m12+iϵ)(q2m22+iϵ)[2πiδp(q+2m12)][2πiδp(q2m22)],

where the subscript “p” of the Dirac-δ functions means that only the contribution of the “proper” root of q+/2=m1/22, for which the temporal component q+/0 is positive, is to be taken. The validity of this substitution rule, termed the Cutkosky rule [4], can be immediately verified within the framework of discontinuity calculus. It is plausible that, for any complex function f(z), even without knowing its analytical expression, the complete analytic continuation and the topological structure of its Riemann surface can be obtained via discontinuity calculus, provided that the specific form of its discontinuity Df(z) is known. This approach can be utilized to perform the analytic continuation of amplitudes with, e.g., triangle and box diagrams [5, 6].

In what follows, we perform a systematic analysis of the discontinuities in the explicit expressions for G(s) using dimensional regularization. In the dimensional regularization scheme, the scalar two-point one-loop integral G(s) in Eq. (8) can be expressed in the following form (see, e.g., [7, 8]):

G(s)=1(4π)2[a(μ)+log(m1m2μ2)+Δslogm1m2+8πL(s)],

where a(μ) is a subtraction constant, and Δ=m12m22. The function L(s) is given by

L(s)=ρ(s)φ(s),

where)

ρ(s)=116πs(ss+)(ss),φ(s)=log[cs16πsρ(s)]log[cs+16πsρ(s)],

with s+=(m1+m2)2 and s=(m1m2)2 being the threshold and pseudo-threshold, respectively, and c=(s++s)/2=m12+m22.

Using (R1) and (R2) and Eq. (7), the action of the discontinuity operator D on the Green’s function G(s) is)

DG(s)=Δ8πilogm1m2δ(s)+12πDL(s).

Then, according to (R3), one has

DL(s)=Dρ(s)φ(s)+ρ(s)Dφ(s)Dρ(s)Dφ(s).

Subsequently, we only need to calculate the discontinuities of ρ(s) and φ(s), respectively. For ρ(s), based on (R1), (R3), and (R4) and Eqs. (5) and (7), one can obtain

Dρ(s)=2ρ(s)θ[(ss+)(ss)]Δ8iδ(s).

The other part φ(s) is the difference between two logarithmic functions. For the first logarithmic function, according to (R1)–(R4) and Eqs. (6) and (13), one can obtain

Dlog[cs16πsρ(s)]=φ(s)θ[(ss+)(ss)]+2πiθ(ss+),

where the relation θ[s±16πsρ(s)c]=θ(ss+) is utilized. On the other hand, since

log[cs16πsρ(s)]+log[cs+16πsρ(s)]=log(c2Δ2)

is a constant, it follows that

Dlog[cs+16πsρ(s)]=Dlog[cs16πsρ(s)].

Consequently, one has

Dφ(s)=2φ(s)θ[(ss+)(ss)]+4πiθ(ss+).

Then, substituting Eqs. (13, 15) into Eq. (12), one obtains

DL(s)=4πiρ(s)θ(ss+)Δ4ilogm1m2δ(s).

Further substituting the above result into Eq. (11) yields

DG(s)=2iρ(s)θ(ss+).

Equation (16) means that G(s) has a branch cut, denoted as Cut-1, in the complex s-plane, starting from the threshold s+ and extending to positive infinity; see the red line in Fig. 1. The continuation kernel of G(s) along Cut-1 is K1G(s)=2iρ(s).

The Riemann sheet corresponding to G(s) is denoted as RS{G(s)}-I, which is called the physical sheet. All other Riemann sheets are called unphysical sheets. The unphysical sheet (lower half-plane) connected to the physical sheet (upper half-plane) along Cut-1 is denoted as RS{G(s)}-II+. The corresponding Green’s function GII+(s) can be obtained by applying the continuation generator T1=1K1 to G(s):

GII+(s)=T1G(s)=G(s)+2iρ(s).

Next, let us examine the discontinuity of GII+(s). Combining Eq. (13) with Eq. (16), one obtains

D[GII+(s)]=DG(s)+2iDρ(s)=2iρ(s)θ(ss+)+4iρ(s)θ(s_s)Δ4δ(s),

where the relation θ[(ss+)(ss)]=θ(ss+)+θ(s_s) has been used.

The above results indicate that there are two cuts (and a pole at s=0) on the Riemann sheet RS{G(s)}-II+. One is Cut-1, with the corresponding continuation kernel K1GII+(s)=2iρ(s). The other, denoted as Cut-2, starts from the pseudo-threshold s and extends toward negative infinity; see the blue line in Fig. 1. Notice that Cut-2 does not appear on the physical sheet. Therefore, a new continuation kernel K2 needs to be introduced, which satisfies K2G(s)=0 and K2GII+(s)=4iρ(s). The analytic continuations of GII+(s) along Cut-1 and Cut-2 are given by the actions of the continuation generators T1 and T2=1K2, respectively:

T1[GII+(s)]=GII+(s)2iρ(s)=G(s),T2[GII+(s)]=GII+(s)4iρ(s)=GII(s),

where we have defined GII(s)G(s)2iρ(s), and the corresponding Riemann sheet is denoted as RS{G(s)}-II. These results demonstrate that RS{G(s)}-II+ (upper half-plane) is connected to the physical sheet (lower half-plane) along Cut-1; simultaneously, RS{G(s)}-II+ (upper half-plane) is connected to RS{G(s)}-II (lower half-plane) along Cut-2.

Furthermore, the discontinuity of GII(s) can be analyzed and its analytic continuation carried out. Since Eq. (19) contains only G(s) and ρ(s), it follows from Eqs. (13, 16) that further analytic continuation will not introduce new continuation kernels. The actions of the continuation generators T1 and T2 on G(s) and ρ(s) are summarized as follows:

T1G(s)=G(s)+2iρ(s),T1ρ(s)=ρ(s),T2G(s)=G(s),T2ρ(s)=ρ(s).

From Eq. (20), the following equations are obtained:

(mN):(T1T2)mG(s)=G(s)+2miρ(s),(T2T1)mG(s)=G(s)2miρ(s).

Based on the above discussions, it can be seen that the analytic continuation of G(s) will generate countably infinite Riemann sheets RS{G(s)}-N±. The Green’s functions on the corresponding Riemann sheets can be defined as follows (N=I,II,III,):

GN±(s)G(s)±2i(N1)ρ(s).

In particular, GI±(s)G(s). These Riemann sheets are connected to each other in the following ways:

 (i) The upper (lower) half-plane of the Riemann sheet RS{G(s)}-N is connected to the lower (upper) half-plane of RS{G(s)}-(N+1)+ along Cut-1;

 (ii) The upper (lower) half-plane of the Riemann sheet RS{G(s)}-N+ is connected to the lower (upper) half-plane of the Riemann sheet RS{G(s)}-N along Cut-2.

These connection patterns can be summarized by the following algebraic expressions:

T1GN(s)=G(N+1)+(s),T1G(N+1)+(s)=GN(s),T2GN+(s)=GN(s),T2GN(s)=GN+(s).

An intuitive representation of the connection patterns among the Riemann sheets RS{G(s)}-N± is shown in Fig. 1. In summary, the structure of the Riemann surface RS{G(s)} can be expressed as follows:

RS{G(s)}=C×{T1,T2|T121,T221}C×Z.

To more intuitively understand the topological structure of the Riemann surface RS{G(s)}, let us extend the domain of G(s) from the complex plane C to the Riemann sphere C^. The Riemann surface obtained through analytic continuation on the Riemann sphere C^ is denoted as RS¯{G(s)}. The topological structure of RS¯{G(s)} can be visualized in the leftmost panel of Fig. 2. In Fig. 2, the red solid line represents Cut-1 on the physical sheet, i.e., the right-hand cut; the orange solid and dashed lines represent Cut-1 on unphysical sheets; the blue solid and dashed lines represent Cut-2 on unphysical sheets. The black dots indicate the branch points of the Green’s functions. It can be clearly observed that the Riemann sheet RS{G(s)}-I is connected to RS{G(s)}-II+ via Cut-1 (the red line). Similarly, RS{G(s)}-II+ is connected to RS{G(s)}-II via Cut-2 (the blue line), while RS{G(s)}-II is connected to RS{G(s)}-III+ via Cut-1 (the orange line). Starting from the physical sheet, crossing Cut-1 and Cut-2 sequentially leads to successive transitions into new unphysical sheets, and this process can continue indefinitely. The Riemann surface RS¯{G(s)} resembles a snail shell composed of individual ridges. Each ridge represents a Riemann sheet, the connections between adjacent ridges correspond to the cuts, and the central point where all connections converge corresponds to the point at infinity.

Before closing this section, we analyze a key special case: the Riemann surface structure of the Green’s function G(s) when analytic continuation across Cut-2 is neglected. From an algebraic perspective, this scenario is equivalent to discarding the continuation generator T2. This results in

RS{G(s)}T2=C×{T1T121}C×Z2.

From a topological perspective, this scenario is equivalent to cutting the snail shell in the leftmost panel of Fig. 2 along Cut-2 (the blue lines). In this case, the structure of the Riemann surface is shown in the middle panel of Fig. 2. The Riemann surface RS¯{G(s)} transforms from a snail shell into a flower with countably infinite petals. These petals are connected exclusively through the receptacle (the point at infinity), with each petal comprising a pair of Riemann spheres RS¯{G(s)}-N and RS¯{G(s)}-(N+1)+. Our analysis focuses on the petal structure encompassing the physical sheet RS¯{G(s)}-I. This branch can be continuously deformed into a sphere while preserving the topological structure (as shown in the rightmost panel of Fig. 2). We will restrict our attention to this branch in the subsequent analysis. In summary, when neglecting the analytic continuation along Cut-2, the Riemann surface RS¯{G(s)} is diffeomorphic to the two-dimensional sphere S2:

RS¯{G(s)}T2S2.

4 Riemann surface of partial-wave scattering matrix

According to unitarity, the two-body partial-wave T-matrix T(s) for two-body coupled-channel scattering) satisfies the optical theorem as follows):

DT(s)=ia=1ncT(s)ϱa(s)T(s)θ(ssa),

where the discontinuity of the matrix T(s) is defined as [DT(s)]abD[Tab(s)]. The matrix indices of T(s) enumerate all two-body coupled channels (totaling nc channels), ordered by increasing threshold energies. The matrix ϱa(s)2Iaρa(s), where (Ia)bc=δabδac, δab is the Kronecker delta, and ρa(s) and sa are the two-body phase space factor and the threshold for channel a, respectively. If the elements of T(s) are real analytic functions, then

T(s+i0+)=T(si0+)=T(s)DT(s).

Substituting the above result into Eq. (23) and comparing it with Eq. (1), one obtains

DT1(s)=ia=1ncϱa(s)θ(ssa).

Recalling Eq. (16), if we introduce the Green’s function matrix as follows:

G(s)diag{G1(s),G2(s),,Gnc(s)},

where Ga(s) is the Green’s function of channel a, then DT1(s)=DG(s). This relation directly leads to the following result:

T(s)=[h(s)+G(s)]1,

where the matrix h(s) satisfies Dh(s)=0.

To determine the Riemann surface structure RS{T(s)}, we first compute the continuation kernels KiT(s), from which we construct the continuation generators Ti=1Ki, and the analytic continuation TiT(s) is obtained. Then, by repeatedly carrying out the analytic continuation of TiT(s), one can obtain the structural information of the Riemann surface RS{T(s)}. Applying the continuation generator Ti to the T-matrix Eq. (25) via the discontinuity rule (R4) yields

TiT(s)=[h(s)+TiG(s)]1.

This result indicates that the analytic continuation of T(s) is completely determined by that of G(s), i.e., RS{T(s)}RS{G(s)}. Therefore, the analysis reduces to studying the analytic continuation of G(s). Equation (24) reveals that the Heaviside θ functions encoding branch cuts in DG(s) exhibit pairwise overlaps. We split each θ function into a sum of boxcar functions, defined as θ[x,y](s)=1 for s[x,y] and 0 otherwise

θ(ssa)=b=ancθ[sb,sb+1](s)b=ancθb(s),

with snc+1 set to +. Substituting the above equation into Eq. (24), exchanging the order of summation, and after rearrangement, one obtains

DG(s)=ia=1ncθa(s)b=1aϱb(s).

From the above equation, the continuation kernel of G(s) along Cut-1 is

K1aG(s)=ib=1aϱb(s),a=1,,nc.

The continuation generator T1a is then defined as T1a1K1a, and the analytic continuation of G(s) along Cut-1 is

T1aG(s)=G(s)+ib=1aϱb(s).

Next, let us consider the analytic continuation of T1aG(s). For simplicity, we assume that the maximum pseudo-threshold across all channels is less than the minimum threshold. According to Eqs. (13) and (27), one can obtain

D[T1aG(s)]=ib=1ncθb(s)[c=1min{a,b}ϱc(s)c>min{a,b}bϱc(s)]+Cut2terms14δ(s)b=1aIbΔb,

where the “Cut-2 terms” in the above formula represent the discontinuity along Cut-2, and Δb=mb12mb22. In the main text, we will not consider the analytic continuation of T1aG(s) along Cut-2, which is not of much interest for physical applications. Interested readers may refer to Appendix C. From Eq. (28), the continuation kernel of T1aG(s) along Cut-1 is

K1b[T1aG(s)]=i[c=1min{a,b}ϱc(s)c>min{a,b}bϱc(s)].

Then, the analytic continuation of T1aG(s) along Cut-1 is

T1b[T1aG(s)]=G(s)+ic>min{a,b}max{a,b}ϱc(s).

From Eq. (29), for any a and b, one has the following relations:

T1b[T1aG(s)]=T1a[T1bG(s)],T1a[T1aG(s)]=G(s).

Thus, the structure of the Riemann surface RS{G(s)} is as follows:

RS{G(s)}=C×{T11,,T1nc|T1a21,T1aT1bT1bT1a(foranya,b)}C×Z2nc.

The above results indicate that, without considering the analytic continuation along Cut-2, the Riemann surface RS{G(s)} consists of 2nc Riemann sheets. The Riemann sheet on which G(s) itself lies is designated as the physical sheet. Among all these Riemann sheets, there are nc Riemann sheets that are directly connected to the physical sheet. The specific forms of the Green’s functions on these sheets are given by Eq. (27), and these nc Riemann sheets are referred to as the first-order unphysical sheets, which are closest to the physical sheet. Additionally, there are Cnc2 Riemann sheets connected to the first-order unphysical sheets. The specific forms of the Green’s functions on these sheets are given by Eq. (29), and these Riemann sheets are called the second-order unphysical sheets, which are farther from the physical sheet. By analogy, there are Cncm m-th order unphysical sheets, and the Green’s functions on these sheets can be obtained by acting with m different continuation generators Ta on G(s). The higher the order m, the farther the sheet is from the physical sheet.

To provide a more intuitive understanding of the above conclusion, we present schematic diagrams of the structure of the Riemann surface RS{G(s)} for nc=2 and 3 in Fig. 3). In Fig. 3, the red lines represent Cut-1 on the physical sheet, while the orange lines represent Cut-1 on the unphysical sheets. Cut-2 has been omitted for simplicity. From Fig. 3, one can observe that in the 2-channel case, there are two first-order unphysical sheets (T11G(s) and T12G(s)) directly connected to the physical sheet, as well as one second-order unphysical sheet (T11T12G(s)). In the 3-channel case, there are three first-order unphysical sheets (T11G(s), T12G(s), and T13G(s)) directly connected to the physical sheet, three second-order unphysical sheets (T11T12G(s), T12T13G(s), and T11T13G(s)), and one third-order unphysical sheet (T11T12T13G(s)). Clearly, the higher the order of the unphysical sheet, the farther it is from the physical sheet.

Moreover, it is evident that all the Riemann sheets in the 3-channel case form a ring-like structure, which is fundamentally different from the topology observed in the 2-channel case. This distinction demonstrates that the Riemann surface RS{G(s)} exhibits different genera in the 2-channel and 3-channel scenarios. A detailed discussion of this aspect will follow in Section 5.

Poles on different Riemann sheets have distinct effects on the invariant mass distribution within the physical region. Typical resonances correspond to poles on first-order unphysical sheets, with the real part of the pole positioned between the branch points that are the endpoints of the red lines in Fig. 3. Such poles are generally closer to the physical sheet and exert a more significant influence on the line shape of the invariant mass distribution within the physical region. If the real part of a pole on the first-order unphysical sheet is not between those branch points represented by the black dots in Fig. 3, the pole can only reach the physical region by circling around one threshold, leading to a threshold cusp in the invariant mass distribution. Similar behavior occurs for all poles on higher-order unphysical sheets. For more detailed discussions on how poles on different Riemann sheets affect the invariant mass line shape, readers are referred to, e.g., Refs. [1113].

Another commonly used labeling scheme exists. To this end, a new set of continuation generators T1a(a=1,2,,nc) is defined,

T1aT1(a1)T1a,

where we stipulate that T10=1. According to Eq. (29), the following equations are obtained,

T1aG(s)=G(s)+iϱa(s),T1a[T1bG(s)]=T1b[T1aG(s)],T1a[T1aG(s)]=G(s).

The physical Riemann sheet is conventionally labeled by the signature “+++” (nc plus signs), where nc denotes the number of scattering channels. For unphysical sheets corresponding to T1aG(s), we flip the a-th positive sign in the physical sheet labeling, e.g., representing the sheet for T11G(s) as “++”. For sheets associated with T1aT1bG(s)(ab), two corresponding signs are flipped, for instance, representing the sheet for T11T12G(s) as “++”. By analogy, for the Riemann sheet corresponding to the Green’s function obtained after applying m different continuation generators T1a to G(s), the corresponding m plus signs in the label of the physical sheet are changed to minus signs.

In this way, all 2nc Riemann sheets can be labeled, and the expressions of the corresponding Green’s functions are given by

G(k)(s)G(s)+ia=1nckaϱa(s),

where k=(k1,k2,,knc)Z2nc, and the Green’s function on the physical sheet is G(s)=G(0)(s).

In summary, we have derived the expressions for G(s) on each Riemann sheet after analytic continuation, as well as the structural information of the Riemann surface RS{G(s)}. Subsequently, based on Eq. (26), one can obtain the expressions for the partial-wave scattering matrix T(s) on each Riemann sheet after analytic continuation:

T(k)(s)=[h(s)+G(k)(s)]1,

together with the structural information of the Riemann surface RS{T(s)}C×Z2nc. This equation can also be written equivalently as

T(k)(s)=T(s)[1+ϱ(k)(s)T(s)]1,

where ϱ(k)(s)=G(k)(s)G(s) is defined in Eq. (33).

5 Uniformization of the partial-wave scattering matrix

In this section, let us briefly discuss the uniformization problem of the partial-wave scattering matrix T(s)—that is, how to map the 2nc Riemann sheets involved in the Riemann surface RS{T(s)} to the interior of the Riemann sphere C^ through a conformal mapping. To this end, one needs to discuss the topological structure of RS¯{T(s)} to obtain the genus of the Riemann surface RS¯{T(s)}. We denote the Riemann surface RS¯{T(s)} as RSnc for brevity, and its genus as gnc.

First, according to Eq. (22), one has RS1S2, and thus g1=0. The structure of RS2 can be obtained through the following three steps:

 ● Cutting: On RS1, start from the threshold s2 on the two Riemann sheets labeled “+” (physical sheet) and “” (unphysical sheet), and make a cut along the real axis.

 ● Duplicating: Add a plus sign “+” to the label of each Riemann sheet on RS1. Then, duplicate RS1 and flip the labels of all Riemann sheets on the duplicate RS1.

 ● Gluing: Glue RS1 and RS1 together according to the connection pattern specified by the analytic continuation to construct the higher-order sheet RS2.

An intuitive representation of these steps is shown in Fig. 4. As illustrated therein, RS2S2 and g2=0. By repeating the cutting, duplication, and gluing procedure, one can generate the higher-order Riemann surface RSnc+1 from the Riemann surface RSnc. The process of obtaining RS3 from RS2 is also depicted in Fig. 4. From Fig. 4, one can observe that RS3 is a torus, RS3S1×S1, and thus g3=1.

Based on this iterative construction process for RSnc, one can derive a recurrence relation for the genera. The genus gnc can be interpreted as the number of holes on the Riemann surface RSnc. When constructing RSnc+1, the total number of holes arises from two sources:

 (i) During duplication, the number of holes on RSnc is doubled. This means that each existing hole on RSnc gives rise to two corresponding holes on RSnc+1.

 (ii) During the cutting procedure, 2nc1 cuts are introduced on RSnc. These cuts contribute 2nc11 additional holes during the subsequent gluing process. This contribution accounts for the newly created holes due to the separation and reconnection of the surface.

Combining both contributions, one has

gnc+1=2gnc+2nc11.

From the above recurrence relation and the initial condition g1=0, one can derive the following general formula for the genus:

gnc=(nc3)2nc2+1.

Our approach concisely reproduces the result in Ref. [14] for the genus formula of the Riemann surface RSnc, which will play a pivotal role in the uniformization of the scattering matrix T(s).

Through the above analysis, one can reduce the uniformization problem of T(s) to the following mathematical question: Does there exist a conformal mapping that can map a compact Riemann surface of genus gnc to the Riemann sphere C^ or its subdomain? If such a mapping exists, what are the characteristics of this conformal mapping?

This question is answered by the uniformization theorem (see, e.g., Ref. [15]) in complex analysis: Every simply connected Riemann surface is conformally equivalent to one of the three canonical Riemann surfaces: the open unit disk D, the complex plane C, or the Riemann sphere C^. This theorem generalizes the Riemann mapping theorem from simply connected open subsets of the complex plane to arbitrary simply connected Riemann surfaces. In particular, for any closed orientable Riemann surface, the corresponding universal cover is determined as follows:

 ● Genus g=0: The Riemann sphere C^ serves as the universal cover, and the corresponding conformal mapping is given by rational functions.

 ● Genus g=1: The complex plane C serves as the universal cover, and the corresponding conformal mapping is given by elliptic functions.

 ● Genus g2: The unit disk D serves as the universal cover, and the corresponding conformal mapping is given by automorphic functions.

Therefore, according to Eq. (34), for the two-channel case, we have g2=0, and the four Riemann sheets of RS{T(s)} can be mapped onto the Riemann sphere C^ via rational functions. For the specific construction of such a mapping, see Ref. [16]. For the three-channel case, we have g3=1, and the eight Riemann sheets of RS{T(s)} can be mapped onto the complex plane C via elliptic functions with two periods. The specific construction of such a mapping was recently carried out in Ref. [10]. For cases where the number of channels satisfies nc>3, we have gnc5, and the 2nc Riemann sheets of RS{T(s)} can be mapped onto the unit disk D via automorphic functions.

6 Summary

In this work, we have proposed a novel method, called the discontinuity calculus, for calculating the analytic continuation and Riemann surface structure of complex functions. This method has been applied to coupled-channel problems in two-body scattering, providing a systematic analysis of the analytic continuation and Riemann surface structure of two-body coupled-channel scattering matrices. Furthermore, we have rederived the genus formula for the Riemann surfaces associated with two-body coupled-channel scattering matrices and established a connection with the uniformization theorem in complex analysis. Our findings are consistent with the established uniformization mapping constructions for two- and three-channel systems, as demonstrated in Refs. [10, 16].

The discontinuity calculus introduced here can also be extended to calculate discontinuities and perform analytic continuation of multivariable complex functions. This extension is applicable for analyzing the analytic continuation and Riemann surface structure of scattering matrices in multi-body (three-body and beyond) coupled-channel scattering problems. Such an extension may provide new insights for analyzing coupled-channel problems involving three-body effects.

7 Appendix A: Alternative definition of the discontinuity calculus

Define the operator H1D. The four properties (R1)–(R4) satisfied by the discontinuity operator D introduced in Section 2 can be equivalently expressed as the following four properties satisfied by the operator H:

(i) Preservation of holomorphy

Hh(z)=h(z).

(ii) Linearity

H[α1f1(z)+α2f2(z)]=α1Hf1(z)+α2Hf2(z).

(iii) Associative homomorphism

H[f1(z)f2(z)]=Hf1(z)Hf2(z).

(iv) Chain rule

HF[f(z)]=HF[ω]|ω=Hf(z).

8 Appendix B: Direct calculation of discontinuities in the two-point Green’s function

The two-point Green’s function G(s) is defined by Eq. (8), which can be rewritten as

G(s)=R4id4q(2π)41D1+D2D3D4+,

where

D1±q0+E/2ω1±iϵ,D2±q0+E/2+ω1±iϵ,D3±q0E/2+ω2±iϵ,D4±q0E/2ω2±iϵ.

The calculation is performed in the center-of-mass frame by setting p=(E,0), where ωi=|q|2+mi2(i=1,2) denotes the on-shell energy of particle-i, and E=s represents the total energy.

We now consider the discontinuity of G(s) with respect to E:

DEG(s)=R4id4q(2π)4DE(1D1+D2D3D4+).

According to (R3), we need to compute the discontinuities of all four terms in the integrand. By convention, the physical region corresponds to the upper edge of the real axis in the complex E-plane. Using Eq. (7), we obtain

DE(1D1±)=2πiδ(q0+E/2ω1)2πiδ1,DE(1D2±)=2πiδ(q0+E/2+ω1)2πiδ2,DE(1D3±)=+2πiδ(q0E/2+ω2)+2πiδ3,DE(1D4±)=+2πiδ(q0E/2ω2)+2πiδ4.

Since the small imaginary component ±iϵ in the argument of the Dirac delta function does not affect subsequent analyses, we have omitted it in the results above for simplicity. Substituting the above results into Eq. (B2) yields

DEG(s)=R4id4q(2π)4[2πi(δ1D2D3D4+δ2D1+D3D4++δ3D1+D2D4++δ4D1+D2D3)(2πi)2(δ1δ2D3D4+δ1δ3D2D4+δ1δ4D2D3δ2δ3D1+D4+δ2δ4D1+D3+δ3δ4D1+D2)+(2πi)3(δ1δ2δ3D4++δ1δ2δ4D3δ1δ3δ4D2δ2δ3δ4D1+)(2πi)4δ1δ2δ3δ4].

In this work, we restrict our analysis to scenarios with non-zero masses m1,m2>0, ensuring that the on-shell energies ω1,ω2>0. This mass condition leads to the vanishing products δ1δ2=δ3δ4=0. Substituting this condition into Eq. (B3), we obtain

DEG(s)=R4id4q(2π)4[2πi(δ1D2D3D4+δ2D1+D3D4++δ3D1+D2D4++δ4D1+D2D3)+(2πi)2(δ1δ3D2D4++δ1δ4D2D3+δ2δ3D1+D4++δ2δ4D1+D3)].

The first line on the right-hand side contains terms with a single Dirac-δ function, referred to as single-pole terms, while the second line comprises terms with two Dirac-δ functions, referred to as double-pole terms.

We first analyze the single-pole terms. For these terms, branch points that contribute to the discontinuity can emerge in the integral only when the pole positions coincide with the endpoints of the complex q0-plane integration contour. In our case, however, the integration path spans the entire real axis, which can be equivalently viewed as a closed contour that closes at infinity, and thus inherently lacks endpoints. Consequently, the single-pole terms do not contribute to the discontinuity.

Next, we investigate the contributions from double-pole terms. For clarity, we henceforth denote the relevant poles directly as δi(i=1,2,3,4). Consider first a generalized integral of the following form:

I(s,n)C(n)×R3id4q(2π)41D1+D2+D3+D4+,

where C(n) represents the integration contour in the complex q0-plane, with the vector n=(n1,n2,n3,n4)Z4 labeling the homotopy equivalence classes of closed contours. Specifically, for n=0, the contour C(0) coincides with the entire real axis. For n0, the contour C(n) comprises not only the real axis but also closed loops encircling each pole δi with winding number ni, where ni>0 indicates counterclockwise encirclements and ni<0 indicates clockwise ones.

From the analysis above, we conclude that only the double-pole terms contribute to the discontinuity of the integral I(s,n). For a specific double-pole term, its contribution depends on whether the integration contour is pinched between the two corresponding poles. In general, we have

DEI(s,n)=C(n)×R3id4q(2π)4[(2πi)2(δ1δ3D2+D4++δ1δ4D2+D3++δ2δ3D1+D4++δ2δ4D1+D3+)]=R4id4q(2π)4{(2πi)2[c13(n)δ1δ3D2+D4++c14(n)δ1δ4D2+D3++c23(n)δ2δ3D1+D4++c42(n)δ2δ4D1+D3+]},

where cij(n)(ij{13,14,23,42}) are integer-valued functions of the homotopy class n, explicitly given by cij(n)=ninj. For an intuitive explanation of this result, we refer to Fig. A1.

Crucially, the two-point Green’s function G(s) in Eq. (B1) specifies that two of the four poles (δ2 and δ3) are located on the upper edge of the complex q0-plane integration contour. By deforming the contour to relocate all poles to the lower edge of the integration contour, we acquire an additional integration over closed loops encircling δ2 and δ3 with winding number +1. Thus, by setting nG=(0,1,1,0), we identify G(s)=I(s,nG). Substitution into Eq. (B4) yields

DEG(s)=DEI(s,nG)=R4id4q(2π)4[(2πi)2(δ1δ3D2+D4+δ2δ4D1+D3+)]=R4id4q(2π)4(2πi)24ω1ω2[δ(Eω1ω2)δ(q0E/2+ω2)+δ(E+ω1+ω2)δ(q0E/2ω2)]=2iρ(E2)[θ(Em1m2)θ(Em1m2)]=2iρ(s)θ(ss+).

This result is completely consistent with the discontinuity formula derived in Eq. (16).

9 Appendix C: Complete analytic continuation of the Green’s function matrix

The “Cut-2 terms” in Eq. (28) are given by

Cut2terms=ib=1ncθb(s)1cap(c)bϱc(s),

where θb(s)θ[sb,sb+1](s), sa (a=1,,nc) are the pseudo-thresholds arranged in descending order, p(c) denotes the order of the pseudo-threshold of channel c, and we set snc+1 to .

We have discussed the analytic continuation of T1aG(s) along Cut-1 in the main text, as shown in Eq. (29). Besides Cut-1, T1aG(s) has another cut, Cut-2, which does not exist on the physical sheet but exists on all other sheets. Thus, one needs to introduce new continuation kernels K2a (a=1,,nc), which satisfy the following equations:

K2aG(s)=0,K2b[T1aG(s)]=i1cap(c)bϱc(s),

and the continuation generators T2a=1K2a:

T2aG(s)=G(s),T2b[T1aG(s)]=G(s)i1cap(c)bϱc(s)+i1cap(c)>bϱc(s).

According to Eq. (27), one can obtain

T1a[T2bG(s)]=T1aG(s)=G(s)+ib=1aϱb(s)T2b[T1aG(s)].

This result indicates that the actions of the continuation generators T1a and T2b on G(s) do not commute. Furthermore, the analytic continuation of T2b[T1aG(s)] across Cut-2 can be evaluated to yield

T2c{T2b[T1aG(s)]}=G(s)+i1dap(d)min{b,c}ϱd(s)i1damin{b,c}<p(d)max{b,c}ϱd(s)+i1dap(d)>max{b,c}ϱd(s).

From this equation, we can derive the following relations:

T2c{T2b[T1aG(s)]}=T2b{T2c[T1aG(s)]},T2b{T2b[T1aG(s)]}=T1aG(s),

which imply the equivalence relations T2aT2bT2bT2a and T2a21.

To clarify the structural information of the Riemann surface RS{G(s)}, we define a new set of continuation generators T2a(a=1,,nc):

T2aT2[p(a)1]T2p(a),

where we stipulate that T20=1. According to Eq. (13), one can obtain

T2aϱa(s)=ϱa(s),T2aϱb(s)=ϱb(s)(ab).

If we replace T2a with T1a from Eq. (31) in the above equations, they still hold.

Furthermore, one can define the raising operator Pa and the lowering operator Pa(a=1,,nc):

PaT1aT2a,PaT2aT1a.

Combining Eqs. (32), (C1), and (C2), one can obtain

P±akG(s)=G(s)±kϱa(s),P±akP±bkG(s)=P±bkP±akG(s)=G(s)±kϱa(s)±kϱb(s),

where k,kZ. These equations imply that the structure of the Riemann surface RS{G(s)} is

RS{G(s)}=C×{P±1,,P±nc|PaPa1,PaPbPbPa(foranya,b)}C×Znc.

In summary, RS{G(s)} consists of countably infinitely many Riemann sheets. The corresponding Green’s function on each Riemann sheet is given by

G(k)(s)G(s)+ia=1nckaϱa(s),

where k=(k1,k2,,knc)Znc.

References

[1]

L. A. Heuser,G. Chanturia,F.-K. Guo,C. Hanhart,M. Hoferichter,B. Kubis, From pole parameters to line shapes and branching ratios, Eur. Phys. J. C 84, 599 (2024), arXiv:

[2]

J. A. Oller, Coupled-channel approach in hadron–hadron scattering, Prog. Part. Nucl. Phys. 110, 103728 (2020), arXiv:

[3]

M. Döring,J. Haidenbauer,M. Mai,T. Sato, Dynamical coupled-channel models for hadron dynamics, arXiv: 2025)

[4]

R. E. Cutkosky , Singularities and discontinuities of Feynman amplitudes, J. Math. Phys. 1, 429 (1960)

[5]

F.-K. Guo,X.-H. Liu,S. Sakai, Threshold cusps and triangle singularities in hadronic reactions, Prog. Part. Nucl. Phys. 112, 103757 (2020), arXiv:

[6]

C.-W. Shen,M.-Y. Duan,J.-J. Wu, Box singularity conditions in box diagrams of decay processes, arXiv: 2025)

[7]

J. A. Oller,E. Oset, N/D description of two meson amplitudes and chiral symmetry, Phys. Rev. D 60, 074023 (1999), arXiv:

[8]

D.-L. Yao,L.-Y. Dai,H.-Q. Zheng,Z.-Y. Zhou, A review on partial-wave dynamics with chiral effective field theory and dispersion relation, Rep. Prog. Phys. 84, 076201 (2021), arXiv:

[9]

J. A. Oller, A Brief Introduction to Dispersion Relations, Springer Briefs in Physics, Springer, 2019

[10]

W. A. Yamada,O. Morimatsu,T. Sato, Analytic map of three-channel S matrix: Generalized uniformization and Mittag−Leffler expansion, Phys. Rev. Lett. 129, 192001 (2022), arXiv:

[11]

X.-K. Dong,F.-K. Guo,B.-S. Zou, Explaining the many threshold structures in the heavy-quark hadron spectrum, Phys. Rev. Lett. 126, 152001 (2021), arXiv:

[12]

Z.-H. Zhang,F.-K. Guo, Classification of coupled-channel near-threshold structures, Phys. Lett. B 863, 139387 (2025), arXiv:

[13]

T. Nishibuchi,T. Hyodo, Eigenstates in coupled-channel scattering amplitude and their effects on spectrum, (2025), arXiv:

[14]

H. A. Weidenmüler , Studies of many-channel scattering, Ann. Phys. 28, 60 (1964)

[15]

W. Schlag, A Course in Complex Analysis and Riemann Surfaces, Graduate Studies in Mathematics, American Mathematical Society, 2014

[16]

M. Kato , Analytical properties of two-channel S-matrix, Ann. Phys. 31, 130 (1965)

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (1417KB)

507

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/