Skip to main content

ORIGINAL RESEARCH article

Front. Phys., 05 November 2019
Sec. Statistical and Computational Physics
Volume 7 - 2019 | https://doi.org/10.3389/fphy.2019.00167

Point-Particle Catalysis

Peter Hayman1,2* Cliff P. Burgess1,2
  • 1Physics & Astronomy, McMaster University, Hamilton, ON, Canada
  • 2Perimeter Institute for Theoretical Physics, Waterloo, ON, Canada

We use the point-particle effective field theory (PPEFT) framework to describe particle-conversion mediated by a flavor-changing coupling to a point-particle. We do this for a toy model of two non-relativistic scalars coupled to the same point-particle, on which there is a flavor-violating coupling. It is found that the point-particle couplings all must be renormalized with respect to a radial cut-off near the origin, and it is an invariant of the flow of the flavor-changing coupling that is directly related to particle-changing cross-sections. At the same time, we find an interesting dependence of those cross-sections on the ratio kout/kin of the outgoing and incoming momenta, which can lead to a 1/kin enhancement in certain regimes. We further connect this model to the case of a single-particle non-self-adjoint (absorptive) PPEFT, as well as to a PPEFT of a single particle coupled to a two-state nucleus. These results could be relevant for future calculations of any more complicated reactions, such as nucleus-induced electron-muon conversions, monopole catalysis of baryon number violation, as well as nuclear transfer reactions.

1. Introduction

It is often the case that physically interesting situations involve a hierarchy of characteristic scales. For instance, solar system dynamics involve a variety of length scales, such as the sizes of the stars and planets involved, as well as the sizes of the orbits. Exploiting such a hierarchy by means of judicious Taylor expansions can greatly simplify otherwise very difficult problems, frequently even providing a handle on seemingly intractable problems. In the realm of quantum field theory, this insight has led to the development of the highly successful effective field theories, which can reduce the complexity of quantum field theories by restricting to parameter subspaces in which an appropriate Taylor expansion can be used to put the theory into a simpler form.

Usually, effective field theories exploit the hierarchy between interaction energies and the masses of some heavy particles to remove those heavy particles from the theory altogether (the quintessential example being Fermi's theory of the Weak interaction, which removes the heavy W and Z bosons) [14]. However, it is often the case that one's interest lies in a sector of the theory that still contains one or two of the heavy particles. For instance, in an atom, a heavy nucleus is present, but for most purposes there is no need to go about computing loops of nucleus-anti-nucleus pairs. Instead, higher energy nuclear dynamics are seen as finite nuclear-size effects. For this reason, an EFT has recently been explored that describes the remnant heavy particles in position-space to exploit the hierarchy of energy scales in a more intuitive expansion in kR, where k is the (small) momentum of the light particle and R is the length-scale of the nuclear structure [57]. This is accomplished in a simple way; the usual effective action is supplemented by a “point-particle” action that involves all possible couplings of the light particle to the worldline of the remnant heavy (point-like) particle (consistent with the symmetries of the low-energy theory). This type of “point-particle” EFT (PPEFT) is conceptually the next best thing to a Fermi type of EFT. While the nuclear dynamics cannot be removed altogether, they are significantly simplified.

The practicality of a PPEFT is two-fold: first it easily permits parameterizing physical quantities in terms of small nuclear properties, since the PPEFT expansion is directly in powers of kR. Second, that parameterization is completely general, and inherently includes all possible interactions, including any potential new physics. Some obvious examples that have been explored are cross-sections and bound-state energies of electrons in terms of nuclear charge radii [7, 8]. In this work, we ask the question: how do the small nuclear properties enter into physical quantities when there are multiple channels of interaction with the point-particle? (We ask this with mind toward eventually describing nuclear transfer reactions, and possibly baryon-monopole reactions which are known to suffer from questions of self-adjointedness of the Hamiltonian similar to those we address in this article [9, 10]).

To answer this question, we consider a simple toy model of two bulk Schrödinger (complex) fields coupled to the same point particle. The most general couplings to the point-particle's worldline yμ(τ) are easily generalized from the single particle species (SP) examples explored in Burgess et al. [5] to the multi-particle (MP) case.

Sb(SP)=-dτ-y˙2[M+hΨ*Ψ+]       Sb(MP)=-dτ-y˙2[M+Ψi*hijΨj+],    (1.1)

where Ψ, Ψi are bulk (complex) scalars, and the flavor index runs over 1 and 2. hij is a matrix of coupling constants that generalizes the single-particle coupling h, and the integral is over the proper time τ of the point-particle (y˙μ:=dyμdτ is the point-particle's 4-velocity). Away from the point-particle, the action is just the usual Schrödinger action for (now) two scalars,

SB(MP)=i=12d4x{ i2(Ψi*tΨiΨitΨi*)  12mi|Ψi|2V(r)|Ψi|2 }.    (1.2)

[V(r) is some bulk potential that may be sourced by the point-particle. In the main text we take it to be an inverse-square potential, since such a potential is highly singular and known to drive interesting behavior in a PPEFT [5]. For the moment it suffices to drop the potential]. If the bulk action (1.2) diagonalizes the momentum it need not diagonalize the brane action (1.1), and it is possible the off-diagonal elements of h (the matrix of hij) can source flavor-violating interactions.

In the center-of-mass reference frame (in the limit of infinite point-particle mass1), the action (1.1) acts as a boundary condition at the origin for the modes of the Ψi fields. However, in general those diverge, and so the action has to be regulated at some finite radius ϵ. The couplings h must then be renormalized to keep physical quantities independent of the regulator, and it turns out that the (low-energy s-wave) cross-section for flavor violation is directly related to an invariant (ϵ3) of the RG-flow of the off-diagonal h12:

σs(12)=4πk2k1ϵ32,    (1.3)

where k1 and k2 are the incoming and outgoing momenta, respectively. For Schrödinger particles, the factor k2/k1=m2/m1 is a constant. However, the same formula holds for spinless relativistic particles, and the ratio k2/k1=(k12+m12-m22)/k12 leads to different qualitative behaviors of the low-energy cross-section depending on how k1 relates to the mass gap m12-m22. If the mass gap is positive, and k12m12-m22, then the cross-section exhibits a 1/k1 enhancement. Both the dependence on ϵ3 and on k1 may prove to be useful in a more complicated calculation, such as in mesic transfer reactions π0 + p → π+ + n (where the neutron and proton are in a nucleus), or possible flavor changing reactions involving new physics, such as μ + Ne + N [11, 12].

The channels of interaction with the point-particle do not have to be different bulk species, however. If, for example, the nucleus carried two accessible energy states, say E = M + Δ/2 and E = M − Δ/2 (where Δ ≪ M is some small excitation energy), then two channels of interaction could be a single bulk particle interacting with each of the nuclear energy eigenstates. In this case the “flavor-violating” cross section is again (1.3), where now k1 and k2 are the incoming and outgoing single-particle momenta, and k2/k1=kout/kin=(kin2±2mΔ)/kin2 with the ± corresponding to the bulk particle impinging on a nucleus in the ground state (−) or the excited state (+), and where m is the mass of the single bulk species (recall we work in the non-recoil limit). On its own, this description is enough for any simple reaction ψ + N → ψ + N*, where an incident non-relativistic particle just knocks a nucleus into a long-lived slightly energized state. Together, the two-species and two-nuclear state models form the building blocks for exploring more complicated processes, such as nuclear transfer reaction, where an incident particle exchanges some constituent particles with a nucleus, and the final state both violates flavor of the bulk species and changes the state of the source nucleus.

Finally, we can also look toward simpler models instead. One may imagine for instance only being interested in tracking one of the bulk-species, say particle 1 (perhaps an apparatus can only detect particles of flavor 1). In that case, the flavor-violating cross-section appears as an absorptive interaction when restricting to the particle 1 subspace of the theory. In this way, our toy model can be seen as a particular unitary completion of a model with a single particle subject to a non-self-adjoint Hamiltonian, as studied in Plestid et al. [13] and frequently used in the form of nuclear optical models [14, 15].

The rest of this paper is organized as follows. First, we briefly recall the salient details of a simple PPEFT for a single bulk species. Then in section 3, we establish the action and classical solutions to a point-particle EFT involving two bulk species, followed by section 4 in which we solve the boundary conditions of the system, and determine how all of the point-particle couplings run. All of this comes together in section 5 where we compute how the point-particle properties relate to physical cross-sections, including the cross-section for flavor-violation. In section 6 we connect the multi-bulk species story to a single particle coupled to a two-state nucleus. Finally, we wrap up in section 7 by restricting to a single-particle subsector of the multi-species model, and realizing the equivalence to the absorptive model of Plestid et al. [13].

2. Point-Particle EFT for a Single Bulk Species

We review the point-particle effective field theory for a single Schrödinger particle in an inverse-square potential, first described in Burgess et al. [5].

In the point-particle effective field theory approach, we exploit the hierarchy of length-scales between the characteristic wavelength of some low-energy particle of mass m (for concreteness, call this some scalar electron) and the scale of some small, almost point-like particle of mass Mm it interacts with (similarly, we will call this a nucleus). For example, in atomic systems, this would be the ratio R/a0 between the size R of a nucleus and the Bohr radius a0 of the atom. For scattering, the small parameter is more directly kR, with k the wavenumber of the incident particle. The way we exploit this hierarchy is to recall that the low energy dynamics of the heavy particle are well approximated by ordinary quantum mechanics, so we imagine only first-quantizing the nucleus. In that case the fully second-quantized electron only couples to the 1-dimensional world-line of the heavy particle. This amounts to writing the action for the electron S = SB + Sb in terms of the usual bulk dynamics2

SB=d4x{i2(Ψ*tΨ-ΨtΨ*)-12m|Ψ|2-V(r)|Ψ|2}    (2.1)

as well as a boundary term consisting of interactions between the electron and the nuclear worldline,

Sb=-dτ-y˙2(M+h|Ψ(y)|2+).    (2.2)

In (2.1), V(r) may be some potential sourced by the point-particle, and the dots represent terms of higher powers in kR. In what follows, we choose V(r)=gr2. On its own, the excessive singularity of the inverse-square potential leads to ambiguities regarding the boundary condition at the origin [16]. Considerations of the self-adjointness of the Hamiltonian are often used to help resolve this difficulty, in particular by means of self-adjoint extensions [1723]. A PPEFT however provides a natural solution to this problem by tying the near-source boundary condition unambiguously to the high-energy physics of the point particle, which may or may not lead to a self-adjoint Hamiltonian for the light degrees of freedom (for an example where probability for the bulk field is lost see [13], and the discussion in chapter 7, below). For a single bulk species this was considered in detail in Burgess et al. [5], and in particular it was shown that the inverse-square potential leads to a non-trivial renormalization of the point-particle coupling h. We briefly recall the results of that calculation below, and in the next section we will ask whether or not similar behavior is seen in the flavor-changing coupling.

Now through (2.2), there are couplings on the world-line yμ(τ) of the nucleus (parameterized by its proper time τ, so that y˙μ:=dyμdτ is the 4-velocity of the nucleus). The first term -y˙2M can be recognized as the usual action for a point-particle [24], while the second term is the lowest-order (in powers of length) coupling between the electron and the nucleus, with the dots representing interactions of higher order in kR. For a spherically symmetric nucleus, the coupling h is a constant.

For simplicity, and to emphasize the value of the point-particle interactions, we work in the limit of infinite nuclear mass, where yμ = (t, 0, 0, 0) is the center-of-mass frame, and τ = t. This amounts to neglecting nuclear recoil, but that can be included by explicitly tracking the dynamics of yμ(τ). Requiring the action to remain stationary with respect to variations of Ψ* that vanish on the boundary yields the usual Schrödinger equation in the bulk

(it+22m-gr2)Ψ=0,    (2.3)

while including variations on the boundary leads to the boundary condition

limϵ04πϵ2ϵΨ=limϵ02mhΨ(ϵ),    (2.4)

which defines ∂ϵ : = ∂r|ϵ, Ψ as the ℓth eigenfunction of angular momentum. However, it is typically the case that one cannot carry out the limit in (2.4)—indeed, in our effective theory, ϵ → 0 is the UV regime we do not claim to understand anyway. The solution to this problem is to replace the point-particle action (2.2) with a boundary action defined on the surface of a ball Bϵ of radius ϵ centered at the world-line of the point-particle (see Appendix A of [5]). In the center of mass frame of reference, this has the form:

SB=-14πBϵd2Ω(M+h|Ψ(ϵ)|2)    (2.5)

and yields the revised boundary condition

4πϵ2ϵΨ=2mhΨ(ϵ).    (2.6)

Here, h has been replaced by a different coupling h to each angular momentum mode of the bulk particle (for more on why this is the case, see [13]). With the limit now gone, the fictitious scale ϵ appears explicitly in the boundary condition for the bulk field, and so runs the risk of appearing in physical quantities. The way around this is to observe that h is not itself a physical quantity, and so must be renormalized with ϵ such that all physical quantities in the boundary condition (2.6) remain ϵ-independent. We explicitly compute this running next.

The bulk equations (2.3) (now defined only outside Bϵ) are solved by

Ψ=e-iEt(C+ψ++C-ψ-)Y,m,    (2.7)

where E is the electron energy, Yℓ, m are the usual spherical harmonics, and the mode functions are

ψ±(ρ):=ρ12(-1±ζ)e-ρ/2M[12(1±ζ),1±ζ;ρ]    (2.8)

which defines k2 : = 2mE, ρ : = 2ikr, and ζ:=(2+1)2-8mg. For simplicity, in this paper we will restrict to the case mg ≤ 1/8 so that ζ is always real. Taking the small limit of (2.8), the boundary condition determines the renormalization-group flow of the coupling h through

λ^=1-C-C+(2ikϵ)-ζ1+C-C+(2ikϵ)-ζ=1+(ϵ/ϵ)-ζ1-(ϵ/ϵ)-ζ,    (2.9)

where λ^:=1ζ(mh/πϵ+1) (we drop the subscript ℓ for convenience), y:=sgn(|λ^|-1) defines a renormalization-group trajectory, and ϵ is an RG-invariant length scale, both determined by the physical quantity

C-C+=-y(2ikϵ)ζ.    (2.10)

Physical quantities like scattering cross-sections and bound-state energies are directly related to the ratio C/C+, and so through (2.10) directly to the quantity ϵ, which is fundamentally a property of the source. The usefulness of the inverse-square potential lies in how it can force non-trivial RG behavior upon the point-particle coupling. For example, the running (2.9) has an “infrared” fixed point of +1 when ϵ/ϵ → ∞, which corresponds to ϵ → 0. For the s-wave in the absence of an inverse-square potential, ζ(ℓ = 0) = 1 and this would be equivalent to vanishing point-particle coupling, but if the strength of the inverse-square potential g ≠ 0, then the fixed point is driven away from a vanishing point-particle coupling.

In the next section, we generalize all of this to a bulk system composed of multiple species of particles (though for concreteness, we specialize to two species). We see how most of the above follows through identically, but the presence of boundary terms that mix flavors adds a new degree of complexity to the problem, introducing a new point-particle coupling which runs differently from (2.9) and opening the door to flavor-changing reactions.

3. Multi-Species Action and Bulk Field Equations

The simplest extension of the basic point-particle action (2.2) to multiple particles is a non-diagonal quadratic one:

Sb(MP)=-dτ-y˙2[M+Ψi*hijΨj+]    (3.1)

where now there are N complex scalar Schroödinger fields Ψi, and summation over the species index is implied. The bulk action is taken to diagonalize the momentum operator, so flavor mixing only happens at the point particle, and the bulk action is simply N copies of (2.1):

SB(MP)=i=1Nd4x{i2(Ψi*tΨi-ΨitΨi*)  -12mi|Ψi|2-V(r)|Ψi|2}.    (3.2)

For concreteness, we will work with only N = 2 species of particles. Our interest is in single particle states, so we restrict to the single-particle sector, for which the Hilbert space is H=H1H2 (where Hi is the Hilbert space for particle i). On this space, the Schrödinger operator it+22mi-gr2 is diagonal. Writing the total wavefunction Ψ=(Ψ1Ψ2), the equations of motion read

[it+22m1-gr200it+22m2-gr2] [Ψ1Ψ2]=0.    (3.3)

The time-dependence is easily solved using separation-of-variables: Ψ=e-iEt(ψ1ψ2). Then we find

[k12+2-gr200k22+2-gr2] [ψ1ψ2]=0,    (3.4)

which defines the wavenumbers ki:=2miE (real for continuum states and imaginary for bound states).

Away from the origin, (3.4) is solved exactly as in the single-species problem, and a natural choice of basis for the solutions is

B={(ψ10),(0ψ2)},    (3.5)

where

ψi=(Ci+ψ+(i;2ikir)+Ci-ψ-(i;2kir))Yimi.    (3.6)

and

ψ±(;ρ):=ρ12(-1±ζi)e-ρ/2M[12(1±ζi),1±ζi;ρ]    (3.7)

with ρ: = 2ikr, and ζi:=(2+1)2-8mig, and Yi, mi the usual spherical harmonics, as in the single-particle example.

The constants Ci± are solved by considering the boundary conditions in the problem, typically finiteness at large- and small-r, but for scattering problems the large-r BC is specific to the setup (since it depends on the presence or otherwise of incident particles). In a PPEFT, the small-r boundary condition is derived from the boundary action, which we describe next.

4. Boundary Conditions

In analogy with the single-particle system, we determine the small-r boundary condition by varying the point-particle action (3.1) directly [including the boundary terms in the variation of the bulk action (3.2)]. The resulting small-r boundary conditions are a simple generalization of (2.6):

4πϵ2ϵΨ=2mhΨ(ϵ),    (4.1)

again using Ψ=e-iEt(ψ1ψ2) and as in section 2, we define ∂ϵ: = ∂r|ϵ (and we drop the angular momentum label on h as above). Here m and h are the mass and point-particle coupling matrices (respectively), so that in components,

4πϵ2ϵψ1-2m1h11ψ1(ϵ)-2m1h12ψ2(ϵ)=0  and    (4.2a)
4πϵ2ϵψ2-2m2h22ψ2(ϵ)-2m2h21ψ1(ϵ)=0.    (4.2b)

Notice that the explicit ϵ-dependence of (4.2) again indicates that the point-particle couplings hij must be renormalized with ϵ whenever Ψ or ∂ϵΨ diverge for small argument, in order for the boundary condition to be compatible with the bulk equations of motion.

The boundary condition then serves two distinct purposes: (i) solving for the integration constants in Ψ tells us how they (and correspondingly physical quantities) depend on the point-source physics, and (ii) isolating for the couplings hij then tells us how exactly each coupling flows with ϵ to ensure the physical integration constants do not. Clearly though, with four possible degrees of freedom in h and four integration constants in Ψ, the two equations in (4.2) are insufficient by themselves. In the next sections we invoke physical arguments to resolve this predicament, and separately tackle both problems (i) and (ii).

4.1. Solving for Integration Constants

The most obvious place to look for additional constraints is at another boundary. In spherical coordinates, this amounts to looking at the asymptotic behavior of Ψ as r → ∞. However, the asymptotic behavior of the system is not unique, and is a very situation-dependent property. Since our interest in this paper is in catalysis of flavor violation, it is most pertinent to study scattering states.

First, focus on scattering Ψ1 → Ψj. In this case, asymptotically we need (see Appendix A.2 for a review of multi-species scattering) ψ1(r)eik1z+f11(θ,ψ)eik1rk1r as usual, and now also ψ2(r)f12(θ,ψ)eik2rk2r. Notice that both boundary conditions and the equations of motion are linear in Ψ, so we may divide through by one integration constant. For incident particle 1, we will choose to divide through by C1+, and we will define

C11:=C1-C1+, and C12:=C2-C1+  (1X scattering),    (4.3)

and eliminating the infalling wave in ψ2 fixes C2+ = RC2− with

R:=-Γ(1-ζ/2)Γ(1+ζ/2)2-2ζe-iπζ    (4.4)

As we will see in the section 5, C11 and C12 are directly related to the physical cross-sections for Ψ1 → Ψ1 and Ψ1 → Ψ2 scattering.

Using the definitions (4.3) and (4.4) in the small-r boundary condition (4.2) yields

4πϵ2ϵ(ψ1++C11ψ1-)-h11(ψ1++C11ψ1-)                                                    -h12C12(Rψ2++ψ2-)=0  and    (4.5a)
4πϵ2C12ϵ(Rψ2++ψ2-)-h22C12(Rψ2++ψ2-)  -h21(ψ1++C11ψ1-)=0,    (4.5b)

Finally, in terms of integration constants the boundary condition (4.5) is now a system of two equations for two unknowns, so using the small-r forms of the bulk modes (3.7),

 ψ±(;ϵ)(2ikϵ)12(-1±ζ)           r|ϵψ±(;ρ)ik(-1±ζ)(2ikϵ)12(-3±ζ),    (4.6)

it is found (see Appendix B for details of the calculation) that the integration constants for this system are related to the source physics through

C11=-(2ik1ϵ)ζ1N^1D^        and  C12=(2ik1ϵ)ζ1/2(2ik2ϵ)ζ2/2m2k2ζ1m1k1ζ2λ^21D^,

where

N^1:=4λ^12λ^21-[λ^11-1][λ^22+1],         andD^:=4λ^12λ^21-[λ^11+1][λ^22+1],    (4.7)

and the following convenient re-definitions have been made:

λ^11:=1ζ1(m1h11πϵ+1),  λ^22:=1ζ2(m2h22πϵ+1),   and,    (4.8)
λ^12:=h12m1m22πϵζ1ζ2,  λ^21:=h21m1m22πϵζ1ζ2.    (4.9)

Of course, the choice to make particle 1 the incident particle was not forced upon us, and with foresight to the next sections, we also compute the quantities involved in scattering Ψ2 → Ψi. Fortunately, this is exactly the 1 ↔ 2 inversion of the 1 → X scattering above, so we can immediately write:

4πϵ2C12ϵ(Rψ2++ψ2-)-     h22C12(Rψ2++ψ2-) -        h21(ψ1++C11ψ1-)=0   and    (4.10a)
4πϵ2ϵ(ψ1++C11ψ1-)-       h11(ψ1++C11ψ1-) -    h12C12(Rψ2++ψ2-)=0,    (4.10b)

where as above we have defined

C22:=C2-C2+, and C21:=C1-C2+  (2X scattering),    (4.11)

now with C1+ = RC1−. Solving for the integration constants similarly yields

C22=-(2ik2ϵ)ζ2N^2D^       and C21=(2ik1ϵ)ζ1/2(2ik2ϵ)ζ2/2m1k1ζ2m2k2ζ1λ^12D^,    (4.12)

where now

N^2:=4λ^12λ^21-[λ^11+1][λ^22-1].    (4.13)

It is important to note that the Cij integration constants are fundamentally different, as they are determined by different asymptotic boundary conditions and correspond to different physics. In section 5 we will see exactly how they relate to the physical cross-sections, but having defined them all separately is already important at the level of renormalizing the point-particle couplings, which we do next.

4.2. Renormalization-Group Flows and Invariants

Next we move on to teasing out of the boundary condition (4.2) exactly how the couplings hij must be renormalized with ϵ to keep physical quantities independent of the regulator. One way to do so would be to differentiate (4.7) and (4.11) with respect to ϵ while holding the (physical) integration constants fixed, and solve the resulting differential equations. This approach turns out to be very difficult however, since the equations are highly coupled and tough to disentangle. Notice however that it is important to have knowledge of both 1 → X and 2 → X scattering to solve for all the elements of h. This is not a coincidence. A unitary system requires a real action, which is enforced by the condition that h is Hermitian. At the same time, a unitary S-matrix for an N-species system has N2 real degrees of freedom, which is the same dimension as an N × N Hermitian matrix. Consequently, connecting the point-particle couplings to physical quantities requires knowledge of the entire S-matrix, and so in our case must involve both 1 → X scattering and 2 → X scattering. Lastly, one final simplification can be made by observing that the phase of h12 can be removed by a field redefinition, so for the special case of a two-species system we only have to deal with a real matrix of point-particle couplings.

The easiest approach to solving for the flows of the couplings hij is to go back to the boundary conditions (4.5) and (4.9) for both 1 → X and 2 → X systems and solve directly for the individual elements of h. This inversion is done in detail in Appendix C, and using the small-r form (4.6), the point-particle couplings must take the following forms as functions of the regulator ϵ.

λ^11=(1-C11(2ik1ϵ)-ζ1)(1+C22(2ik2ϵ)-ζ2)+C21C12(2ik1ϵ)-ζ1(2ik2ϵ)-ζ2(1+C11(2ik1ϵ)-ζ1)(1+C22(2ik2ϵ)-ζ2)-C21C12(2ik1ϵ)-ζ1(2ik2ϵ)-ζ2,    (4.14a)
λ^12=m2ζ1m1ζ2C21(k1k2)-1/2(2ik1ϵ)-ζ1/2(2ik2ϵ)-ζ2/2(1+C11(2ik1ϵ)-ζ1)(1+C22(2ik2ϵ)-ζ2)-C21C12(2ik1ϵ)-ζ1(2ik2ϵ)-ζ2,    (4.14b)
λ^21=m1ζ2m2ζ1C12(k2k1)-1/2(2ik1ϵ)-ζ1/2(2ik2ϵ)-ζ2/2(1+C11(2ik1ϵ)-ζ1)(1+C22(2ik2ϵ)-ζ2)-C21C12(2ik1ϵ)-ζ1(2ik2ϵ)-ζ2,    (4.14c)
λ^22=(1+C11(2ik1ϵ)-ζ1)(1-C22(2ik2ϵ)-ζ2)+C21C12(2ik1ϵ)-ζ1(2ik2ϵ)-ζ2(1+C11(2ik1ϵ)-ζ1)(1+C22(2ik2ϵ)-ζ2)-C21C12(2ik1ϵ)-ζ1(2ik2ϵ)-ζ2,    (4.14d)

using the definitions (4.9).

Equations (4.13) (together with (4.7), (4.11), (2.10), and past work [8]) suggests the integration constants can each be characterized by a unique RG-invariant length-scale. To see how this works, define the scales ϵ1, ϵ2, and ϵ3 by the following relations:

 C11=-y1(2ik1ϵ1)ζ1,      C22=-y2(2ik2ϵ2)ζ2,and  C12=m2k2ζ1m1k1ζ2C21=y3m2k2ζ1m1k1ζ2(2ik1ϵ3)ζ1/2(2ik2ϵ3)ζ2/2,    (4.15)

where yi = ±1 defines a particular class of flow. In terms of these scales, the running equations are significantly simpler:

λ^11=(1+y1(ϵ/ϵ1)-ζ1)(1-y2(ϵ/ϵ2)-ζ2)+(ϵ/ϵ3)-(ζ1+ζ2)(1-y1(ϵ/ϵ1)-ζ1)(1-y2(ϵ/ϵ2)-ζ2)-(ϵ/ϵ3)-(ζ1+ζ2),    (4.16a)
λ^12=λ^21 =y3(ϵ/ϵ3)-(ζ1+ζ2)/2(1-y1(ϵ/ϵ1)-ζ1)(1-y2(ϵ/ϵ2)-ζ2)-(ϵ/ϵ3)-(ζ1+ζ2),    (4.16b)
λ^22=(1-y1(ϵ/ϵ1)-ζ1)(1+y2(ϵ/ϵ2)-ζ1)+(ϵ/ϵ3)-(ζ1+ζ2)(1-y1(ϵ/ϵ1)-ζ1)(1-y2(ϵ/ϵ2)-ζ2)-(ϵ/ϵ3)-(ζ1+ζ2),    (4.16c)

An example of λ^11 and λ^22 is plotted in Figure 1, and an example of λ^12 is plotted in Figure 2. All the couplings flow to fixed points in the ultraviolet (ϵ/ϵi → 0) and the infrared (ϵ/ϵi → ∞). The diagonal couplings λ^11 and λ^22 both flow to −1 in the UV and +1 in the IR, exactly as the single-particle flow does, while the off-diagonal λ^12 flows to vanishing coupling in both the cases. This says something reasonable: the system is perfectly content to live in a world where there is no species mixing, regardless of the existence of the inverse-square potential. However, if the species do mix, then the strength of that mixing depends on the scale it is measured at, with the flow given by (4.15b).

FIGURE 1
www.frontiersin.org

Figure 1. Plot of λ^11 vs ζ ln(ϵ/ϵ1) for ζ1 = ζ2, y1 = y2 = +1, and (ϵ2/ϵ1)ζ=2 and (ϵ3/ϵ1)ζ=0.02. The fascinating second pole arises in certain limits of the ratio of ϵ1 to ϵ2 and ϵ3. In the limit ϵ3 → 0, this reduces to the classic single-particle RG.

FIGURE 2
www.frontiersin.org

Figure 2. Plot of λ^12 vs ζ ln(ϵ/ϵ1) for ζ1 = ζ2, y1 = y2 = y3 = +1, and (ϵ2/ϵ1)ζ=2 and (ϵ3/ϵ1)ζ=0.02.

Interestingly, all three flows share a common denominator, which can always be factorized. When ζ1 = ζ2 = : ζ, the zeroes of the denominator lie at

ϵaζ=12(y2ϵ2ζ+y1ϵ1ζ±(y1ϵ1ζ-y2ϵ2ζ)2+4ϵ32ζ),    (4.17)

so there is at least one asymptote in all the flows as long as y2ϵ2ζ+y1ϵ1ζ>0. Indeed, the only regime where there is no asymptote is where -(y2ϵ2ζ+y1ϵ1ζ)>(y1ϵ1ζ-y2ϵ2ζ)2+4ϵ32ζ>0.

The practicality of this framework lies in how the point-particle couplings [and in particular their RG-invariants (4.14)] inform physical quantities, like scattering cross-sections, and we investigate this next.

5. Scattering and Catalysis of Flavor Violation

To see how the nuclear properties enter into macroscopic quantities, and in particular how the point-particle can induce a violation of bulk flavor-conservation, we proceed to compute the elements of the scattering matrix. In particular, it will be shown that the low-energy s-wave “elastic” (Ψi → Ψi) scattering is as usual independent of incoming momentum, and ϵi plays the role of the scattering length. Meanwhile, the flavor-violating “inelastic” cross-section is uniquely characterized by ϵ3, which can be thought of as an effective scattering length for flavor violation. Moreover, the inelastic cross-section goes as kout/kin. This ratio is a constant for Schrödinger particles, but for Klein-Gordon fields (such as is appropriate for, say, incident pions) the ratio has a dependence on the incoming momentum, and that dependence takes on a variety of qualitatively different forms determined by the size of kin2/|m12-m22|. Of particular interest is the case where the incoming particle's mass is greater than the outgoing particle's mass, and the incoming particle's momentum is small compared to the mass gap, in which case the flavor-violating cross-section is catalyzed, and goes as 1/kin (although such a setup would certainly be hard to realize in practice).

Following the Appendix A.2, we need to identify the scattering matrix elements Sij() in terms of Cij.

5.1. Elastic Scattering (ψi → ψi)

First consider elastic scattering of ψ1. This is the case where the measured particle is the same as the incoming particle, so the asymptotic form of ψ1 contains an incoming plane wave and an outgoing spherical wave:

ψ1Ans(r)C(eik1z+f11(θ,ψ)eik1rr).    (5.1)

Starting from the general form (3.6) for particle 1,

ψ1=(C1+ψ1++C1-ψ1-)Y1,m1,    (5.2)

the large-r limit is (taking the asymptotic limit of the confluent hypergeometric function)

ψ1(r)Aei(k1r-π/2)2ik1r+Be-i(k1r-π/2)2ik1r,    (5.3)

where

A=[Γ(1+12ζ1) 2ζ1C1++Γ(1-12ζ1) 2-ζ1C1-]eiπ/2π, andB=[Γ(1+12ζ1) 2ζ1C1++Γ(1-12ζ1) 2-ζ1e-iπζ1C1-]ei(1+ζ1-)π/2π.    (5.4)

Matching to the asymptotic form (5.1), allows us to identify (see 12.1) the overall normalization

C=(-1)+12π(2+1)[Γ(1+12ζ1)2ζ1C1+  +Γ(1-12ζ1) 2-ζ1e-iπζ1C1-]ei(1+ζ1)π/2,    (5.5)

and the scattering matrix element

 S11()=-AB=[Γ(1+ζ1/2)+Γ(1-ζ1/2)2-2ζ1C11][Γ(1+ζ1/2)+Γ(1-ζ1/2)2-2ζ1e-iπζ1C11]ei(2+1-ζ1)π/2.    (5.6)

Our interest is in the regime where kiϵ is small, so the s-wave is the dominant contribution to the cross-section. In the s-wave, ζ1=ζ1s:=1-8m1g, and the cross-section is (A.9), exactly as in Burgess et al. [6]:

σs(11)=πk12|S11(0)|2=πk12|1-Aeiπζ1s/21-Ae-iπζ1s/2|2    (5.7)

where

A:=e-iπζ1/22-2ζ1C11Γ[1-12ζ1]Γ[1+12ζ1]=y1(k1ϵ12)ζ1Γ[1-12ζ1]Γ[1+12ζ1].    (5.8)

[The second equality uses (4.14) to exchange the integration constants for the RG-invariant ϵ1]. Of particular note is when there is no inverse-square potential, in which case ζ1s = 1 and the cross-section reduces to

σs(11)=4πϵ12  (g=0),    (5.9)

which can be identified as the cross-section for scattering from a 3D δ-function potential (see for example [25] where our ϵ1 corresponds to their g/π). Elastic scattering for the second species ψ2 → ψ2 follows exactly the same procedure and is trivially the 1 ↔ 2 inversion of (5.7) and (5.8).

σs(22)=4πϵ22  (g=0).    (5.10)

5.2. Flavor-Violating Scattering (ψi → ψj, ij)

So much for the ordinary scattering. Finally we compute the flavor-violating ψ1 → ψ2 cross-section, and see the point-particle catalysis in action. This time, there is no incoming flux of the particle to be measured, so the large-r ansatz is simply:

ψ2Ans(r)Cf2(θ,ψ)eik2rr.    (5.11)

where we have scaled out the same normalization factor C, for convenience. The asymptotic form of the general solution for ψ2 is exactly (5.3) subject to 1 ↔ 2, so we can immediately observe the following. First, as was used in deriving the boundary conditions in section 4, B = 0 so that C2+ = RC2−, where as above

R:=-Γ(1-ζ/2)Γ(1+ζ/2)2-2ζe-iπζ.    (5.12)

Following Appendix A.2, this time we identify the inelastic scattering element S12() using (A.15)

S12()=e-iπ/24π(2+1)AC=Γ(1-12ζ2)2π(2+1)C2-C2-ζ2(1-e-iπζ2), =(-1)-+12ζ1+ζ2-1 sin (πζ2/2)eiπ(-ζ1-ζ2-1)/2Γ(1-12ζ2)Γ(1+12ζ1)C121-Ae-iπζ1/2.    (5.13)

where A is defined in (5.8).

Again, our interest is in the small kiϵ regime for which the s-wave dominates. We similarly define ζ2s:=1-8m2g. Then the low-energy cross-section is (A.16),

σs(12)=πk1k2m1m2|S12(0)|2, =4π2ζ1+ζ2k1ζ1k2ζ2k12ζ1ζ2 sin2(πζ2/2)ϵ3ζ1+ζ2|1-Ae-iπζ1/2|2(Γ(1-12ζ2)Γ(1+12ζ1))2.    (5.14)

As in the elastic case, the reverse scattering ψ2 → ψ1 is a simple matter of exchanging 1 ↔ 2 in (5.14). In the absence of an inverse-square potential (ζis = 1), the cross-section (5.14) simplifies significantly.

σs12=4πk2k1ϵ32  (g=0).    (5.15)

On its own, this is an interesting enough result. The flavor-violating cross-section is non-zero only when the point-particle has non-trivial flavor-violating properties, as are encoded in ϵ3. This is also the statement that flavor-violation only occurs if h12 ≠ 0, since it is always true—regardless of the presence of an inverse-square potential—that h12 = 0 only when ϵ3 = 03.

A particularly interesting aspect of (5.15) is the factor of k2/k1. For Schrödinger particles, this is just a constant (ki=2miEk2/k1=m2/m2). In section 6, we treat a Schrödinger particle interacting with a multi-state nucleus, in which case the final and initial momenta do differ non-trivially, but even at the level of two bulk species, more interesting dynamics can be seen just by treating the particles as relativistic Klein-Gordon fields. For relativistic fields, the cross-section takes the same general form as (5.15), except the momenta are relativistic: ki=ω2-mi2, where ω is the energy of the system. Then ω2=ki2+mi2, so that

k2k1=k12+(m12-m22)k1.    (5.16)

The cross-section therefore exhibits different qualitative behavior depending on how the incoming momentum k1 relates to the (squared) mass gap m12-m22. This can be broadly classified by 4 different regimes depending on the sign of m12-m22 and the size of the ratio r:=k12/|m12-m22|.

First, if the mass gap is positive, the only question is to the size of the ratio r. If r ≪ 1, then σs(12)~k1-1. This is the regime where the cross-section sees a low-energy enhancement similar to the well-known enhancement of absorptive cross-sections [26] (more on that in section 6). However, if r ≫ 1, then the cross-section is roughly independent of the incoming momentum altogether. The cross-over between these regimes is plotted in Figure 3. These are indeed reasonable behaviors. If m1 > m2, and the incoming momentum is small compared to the mass difference, then the transition is to a lighter particle traveling faster, which is intuitively a more favorable process—the heavier incident particle has access to a larger phase-space than the lighter incident particle. If the mass difference is small compared to the incoming momentum, then the benefit of transitioning to a particle with a smaller mass is minimal, so the process is no more favorable than no transition. If instead the mass gap is negative, then there is a new regime. If k12<m22-m12 (so if r < 1) there is in fact no scattering. This is certainly reasonable—if the incident particle did not have enough energy to create the rest mass of the second particle, then it cannot scatter into that particle (this is the threshold behavior described by [27, section 144]). If r ≳ 1, the incident momentum is just enough to create the second particle k12=m22-m12+δ, then the cross-section goes as δ/(m22-m12). Since k22=k12+m12-m22=δm22, this is also the statement that the cross-section goes as v2, the (non-relativistic) speed of the final state particle. Finally, if r ≫ 1, the incident momentum greatly exceeds the mass gap and we again see the cross-section behave independently from k1, as before. These momentum-dependences are plotted in Figure 4.

FIGURE 3
www.frontiersin.org

Figure 3. Plot of the cross-over behavior in the k dependence of the inelastic cross-section for a Klein-Gordon particle conversion in units of Δ/k1, where Δ=m12-m22 and m1 > m2. The full function 1+Δ2/k12 is plotted in blue, and the simple inelastic behavior Δ/k1 is plotted in orange. Notice the overlap for small Δ/k1, and the strong enhancement for large momenta.

FIGURE 4
www.frontiersin.org

Figure 4. Plot of the cross-over behavior in the k dependence of the inelastic cross-section for a Klein-Gordon particle conversion in units of Δ/k1, where Δ=m12-m22 and m2 > m1. The full function 1-Δ2/k12 is plotted in blue, and the simple inelastic behavior Δ/k1-1 is plotted in orange. Notice the threshold cutoff at Δ/k1 = 1, as well as the approximate overlap for small Δ/k1, and the strong suppression for large momenta.

6. Transfer Reactions and Nuclear Structure

In many cases, a reaction with a nucleus can change not only the incident particle, but also the nucleus. This can be the case even when scattering energies are low compared to the mass of the nucleus. For instance, the excitation energy of most real nuclei is on the order of MeV compared to their masses of order GeV [28]. A particularly interesting class of reactions that falls into this category is transfer reactions, where a composite particle (say a neutron) scatters off of a nucleus and exchanges a constituent particle (say a quark) with one of the valence nucleons, so that the outgoing particle is different (perhaps a proton) and so is the nucleus. While this work is not enough to describe a complete transfer reaction, we can make progress toward a complete description, and can at least describe the simpler process ψ + N → ψ + N*, where N is some nucleus and N* is a long-lived excited state of that nucleus. The key observation to make is that there is essentially no difference between a system spanned by {Ψ1 ⊗ |N〉, Ψ2 ⊗ |N〉} with |N〉 some nuclear state, and {Ψ ⊗ |N1〉, Ψ ⊗ |N2〉} where |Ni〉 are distinct nuclear states. The only complexity lies in describing the different nuclear states in a point-particle EFT language.

Here we will sketch out the simplest point-particle action that includes a two-state nucleus coupled to a single bulk field, but a more detailed treatment of a PPEFT for a point-particle with internal degrees of freedom will be available in Zalavái et al. (in preparation). In addition to the single-species action (2.2), introduce an auxiliary grassmann-valued field Ti that satisfies the commutation relations of the generators of 𝔰𝔲(2).

Sb2N=-dτ-y˙2(M+iTiT˙i-iϵijkΔiTjTk+h|Ψ(y(τ))|2   -ϵijkgiTjTk|Ψ(y(τ))|2).    (6.1)

Here, Δi and gi are 3-vector-valued parameters. For convenience, we work in the basis such that Δi = Δδi3. Furthermore, we collect the point-particle couplings involving Ψ as W=W:=h-i2giTi.

Upon quantization, the Ti can be identified with i2σi, the generators of 𝔰𝔲(2), and since they only live on the point-particle's world-line, it is easy to see that they are associated with a two-level nuclear state. Varying the action with respect to y(τ) (and neglecting the subdominant contribution from the W interactions), the nuclear dispersion relation

N|p^N2-M2+Δσz|N=0,    (6.2)

(p^N is the nuclear 4-momentum operator) leads to distinct nuclear states |↑〉 with rest-frame energy E = M + Δ/2 and |↓〉 with rest-frame energy E = M − Δ/2.

The bulk action for the system is exactly the single-particle Schrödinger action (2.1), and so the solutions for Ψ are precisely (2.7) and (2.8). However, the boundary condition (2.6) becomes:

|4πϵ2ϵΨ|Ξ=|WΨ(ϵ)|Ξ,    (6.3)

where |Ξ〉 is an appropriate Fock state, |↑↓〉 are the Fock states consisting of just the nucleus in the state with rest-frame energy E↑↓, and Ψ is interpreted as an operator-valued field. Since energy can now be exchanged between the electron and the nucleus, the individual energies of the electron and the nucleus are no longer good quantum numbers, and a general single-electron Fock state must be a linear combination |Ξ〉 = |Ψ〉|↑〉 + |Ψ〉|↓〉, where |Ψ↑↓〉 has energy ω↑↓ satisfying ω + Δ/2 = ω − Δ/2. Then in terms of the mode-functions Ψ(x)=e-iωψ (satisfying Ψ|Ψ↑↓〉 = Ψ↑↓(x)|Ψ↑↓〉), the boundary condition (6.3) is in components,

4πϵ2ϵψ-Wψ(ϵ)-Wψ(ϵ)=0  and4πϵ2ϵψ-Wψ(ϵ)-Wψ(ϵ)=0,    (6.4)

where we identified the Ti with 12σi so that W=h𝕀2×2+14giσi. The boundary condition (6.4) is now exactly the boundary condition (4.2) with Wh and ψ↑↓ ↔ ϕ1, 2.

Finally, defining E: = ω + Δ/2 = ω − Δ/2, we observe k2=2mω=2m(E-Δ/2) and k2=2mω=2m(E+Δ/2). Choosing ψ ↔ ψ1 and ψ ↔ ψ2, we may then identify k1 = k and k2 = k. At this point, a complete analogy with the two-species system is established, and the results are tabulated into a dictionary relating the two in Table 1. A true transfer reaction is one for which the final state involves both a different species of bulk particle and an altered state of the nucleus, so evidently would be equivalent to a model of 4 bulk species coupled to a single-state point-particle. As it stands however, this system is sufficient to describe, say, the low-energy behavior of a neutron that knocks a nucleus into its first excited state.

TABLE 1
www.frontiersin.org

Table 1. The dictionary that maps quantities in a two-bulk-species theory to quantities in a theory of a single bulk-species coupled to a point-particle with two accessible internal degrees of freedom.

7. Single-Particle Subsector

In many cases it is overkill to track all of the possible final state products of a particular interaction. This is especially the case in nuclear physics, where a summation over many unobserved final states is the basis of the highly successful optical model [14, 15]. The price one pays for the convenience of ignoring certain states is the loss of unitarity, and such a non-unitary point-particle EFT was the subject of Plestid et al. [13]. Here we can provide a very simple explicit example of how this non-unitarity can emerge in a subsector of a larger unitary theory. We achieve this correspondence by matching physical quantities, in a procedure that is significantly simpler than e.g., tracing the partition function over the states involving Ψ2 [29, 30].

From Plestid et al. [13], the key to a point-particle inducing a violation of unitarity is allowing the point-particle coupling to be complex (h from section 2). In that case, the running of the coupling is the same as (2.9), except now the constant -yeiα is complex. That is:

λ^c=1-eiα(ϵ/ϵ)-ζ1+eiα(ϵ/ϵ)-ζ,    (7.1)

where λ^c is the now complex coupling. Similarly, the integration constant is analogous to (2.10),

C-C+=(2ikϵ)ζ1-λ^c1+λ^c=(2ikϵ)ζeiα.    (7.2)

At this point, we essentially have everything we need. The ratio of integration constants C-C+ is directly related to the physical quantities in the single-particle problem, so choosing to track either Ψ1 or Ψ2 tells us to equate (7.2) to C11 or C22 (respectively), and from that determine how the RG-invariants and couplings are related. The only obstruction at this point is that (4.14) would at face value suggest that α = and there is no absorptive scattering. The error here is that inelastic scattering is a sub-leading effect4, and to see it at the level of C11, we would need to have computed that ratio of integration constants to sub-leading order in kiϵ. This is not in itself a particularly challenging endeavor, and is done in the Appendix B. The result is that to sub-leading order, one finds (B.22) (choosing to track Ψ1, tracking Ψ2 follows trivially)

C11=-(2ik1ϵ)ζ1N^1D^(1-4Rλ^12λ^21N^1D^(2ik2ϵ)ζ2) =-y1(2ik1ϵ1)ζ1(1+iδα1).    (7.3)

From the last equality, we use that δα1 ≪ 1 to define eiα1-y1(1+iδα1) such that α1: = + δα1, with n an integer that satisfies y1:=-einπ. Since (7.3) is a perturbative expression, we may use the leading order (in kiϵ) expressions for the couplings in δα1, and simply evaluate them at ϵ = ϵ1. An alternative but more tedious approach would be to substitute the second equality in (7.3) into λ^11 [taking the whole function to sub-leading order in kiϵ as in (C.12)] and solve for δα1 by demanding λ^11 remain real at sub-leading order, as it must. No matter the approach, the result is

δα1=-(2k2ϵ3)ζ1(ϵ3ϵ1)ζ2|R|y1 sin(πζ2/2).    (7.4)

In this way, we have solved for the RG-invariant quantities (and so too the physical quantities) in the single-particle absorptive model in terms of the RG-invariants in the unitary two-species model simply by equating C11 to C/C+. In fact, we can do even better than that. We can determine how the coupling λ^c relates to the various λ^ij couplings. To do so, we simply arrange for N^A/D^ to take the form of (λ^c-1)/(λ^c+1).

N^1D^=4|λ^12|2-[λ^11-1][λ^22+1]4|λ^12|2-[λ^11+1][λ^22+1]=λ^11[λ^22+1]-4|λ^12|2[λ^22+1]-1λ^11[λ^22+1]-4|λ^12|2[λ^22+1]+1.    (7.5)

Evidently λ^c=λ^11-4|λ^12|2/[λ^22+1]. A check on this is to directly compute the combination λ^11-4|λ^12|2/[λ^22+1], in which case one finds it is exactly (7.1) with ϵ and α defined as above. This dictionary between these models is laid out in Table 2.

TABLE 2
www.frontiersin.org

Table 2. The dictionary that maps quantities in a unitary two-species theory to quantities in a non-unitary single-species theory.

Data Availability Statement

All datasets generated for this study are included in the article/supplementary material.

Author Contributions

PH wrote most of the text and derived most of the results. CB provided significant guidance, corrected errors, helped with interpretations, and provided feedback on the manuscript.

Funding

This research was supported in part by funds from the Natural Sciences and Engineering Research Council (NSERC) of Canada. Research at the Perimeter Institute is supported in part by the Government of Canada through Industry Canada, and by the Province of Ontario through the Ministry of Research and Information (MRI).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank László Zalavári, Daniel Ruiz, Markus Rummel, and Ryan Plestid for helpful discussions.

Footnotes

1. ^We neglect here recoil effects, though those can be included by tracking the dynamics of yμ(τ).

2. ^We use a mostly plus metric, and work in units such that ℏ = c = 1.

3. ^As noted in section 4.2, this is in contrast to the behavior of h11 and h22, whose flows indicate that ϵ1, ϵ2 → 0 is only consistent with h11, h22 → 0 when there is no inverse-square potential.

4. ^The way to see this is through the catalysis cross-section (5.15). Absorptive scattering generically scales as a/k for some absorptive scattering length a. The derived cross-section (5.15) identifies a ~ (k2ϵ33 and so is generically a subdominant effect in the point-particle EFT regime.

References

1. Burgess CP. An introduction to effective field theory. Annu Rev Nucl Part Sci. (2007) 57:329–62. doi: 10.1146/annurev.nucl.56.080805.140508

CrossRef Full Text | Google Scholar

2. Burgess CP, Moore GD. The Standard Model: A Primer. Cambridge: Cambridge University Press (2006).

Google Scholar

3. Weinberg S. The Quantum Theory of Fields. Vol. 1. Cambridge: Cambridge University Press (1995).

4. Weinberg S. The Quantum Theory of Fields. Vol. 2. Cambridge: Cambridge University Press (1996).

Google Scholar

5. Burgess CP, Hayman P, Williams M, Zalavári L. Point-particle effective field theory I: classical renormalization and the inverse-square potential. J High Energy Phys. (2017) 2017:106. doi: 10.1007/JHEP04(2017)106

CrossRef Full Text | Google Scholar

6. Burgess CP, Hayman P, Rummel M, Williams M, Zalavári L. Point-particle effective field theory II: relativistic effects and Coulomb/inverse-square competition. J High Energy Phys. (2017) 2017:72. doi: 10.1007/JHEP07(2017)072

CrossRef Full Text | Google Scholar

7. Burgess CP, Hayman P, Rummel M, Zalavári L. Point-particle effective field theory III: relativistic fermions and the Dirac equation. J High Energy Phys. (2017) 2017:7. doi: 10.1007/JHEP09(2017)007

CrossRef Full Text | Google Scholar

8. Burgess CP, Hayman P, Rummel M, Zalavri L. Reduced theoretical error for 4He+ spectroscopy. Phys Rev A. (2018) 98:052510. doi: 10.1103/PhysRevA.98.052510

CrossRef Full Text | Google Scholar

9. Kazama Y, Yang CN, Goldhaber AS. Scattering of a Dirac particle with charge Ze by a fixed magnetic monopole. Phys Rev D. (1977) 15:2287–99. doi: 10.1103/PhysRevD.15.2287

CrossRef Full Text | Google Scholar

10. D'Hoker E, Farhi E. The abelian monopole fermion system and its self-adjoint extensions as a limit of a non-abelian system. Phys Lett B. (1983) 127:360–4.

Google Scholar

11. Pezzullo G. The Mu2e experiment at Fermilab: a search for lepton flavor violation. Nucl Part Phys Proc. (2017) 285-86:3–7. doi: 10.1016/j.nuclphysbps.2017.03.002

CrossRef Full Text | Google Scholar

12. Bonventre R. Searching for muon to electron conversion: the Mu2e experiment at Fermilab. SciPost Phys Proc. (2019) 1:038. doi: 10.21468/SciPostPhysProc.1.038

CrossRef Full Text | Google Scholar

13. Plestid R, Burgess CP, O'Dell DHJ. Fall to the centre in atom traps and point-particle EFT for absorptive systems. J High Energy Phys. (2018) 2018:59. doi: 10.1007/JHEP08(2018)059

CrossRef Full Text | Google Scholar

14. Feshbach H. Nuclear reactions. In: Digital Encyclopedia of Applied Physics. American Cancer Society (2003). Available online at: https://onlinelibrary.wiley.com/doi/abs/10.1002/3527600434.eap277

Google Scholar

15. Dickhoff WH, Charity RJ. Recent developments for the optical model of nuclei. Prog Part Nucl Phys. (2019) 105:252–99. doi: 10.1016/j.ppnp.2018.11.002

CrossRef Full Text | Google Scholar

16. Essin AM, Griffiths DJ. Quantum mechanics of the 1/x2 potential. Am J Phys. (2006) 74:109–17. doi: 10.1119/1.2165248

CrossRef Full Text | Google Scholar

17. Capri AZ. Self-adjointness and spontaneously broken symmetry. Am J Phys. (1977) 45:823–25. doi: 10.1119/1.11055

CrossRef Full Text | Google Scholar

18. Dupré MJ, Goldstein JA, Levy M. The nearest self-adjoint operator. J Chem Phys. (1980) 72:780–1. doi: 10.1063/1.438923

CrossRef Full Text | Google Scholar

19. Bonneau G, Faraut J, Valent G. Self-adjoint extensions of operators and the teaching of quantum mechanics. Am J Phys. (2001) 69:322–31. doi: 10.1119/1.1328351

CrossRef Full Text | Google Scholar

20. Borg JL, Pulé JV. Pauli approximations to the self-adjoint extensions of the Aharonov-Bohm Hamiltonian. J Math Phys. (2003) 44:4385–410. doi: 10.1063/1.1601298

CrossRef Full Text | Google Scholar

21. Araujo VS, Coutinho FaB, Fernando Perez J. Operator domains and self-adjoint operators. Am J Phys. (2004) 72:203–13. doi: 10.1119/1.1624111

CrossRef Full Text | Google Scholar

22. Camblong HE, Epele LN, Fanchiotti H, García Canal CA, Ordónez CR. On the inequivalence of renormalization and self-adjoint extensions for quantum singular interactions. Phys Lett A. (2007) 364:458–64. doi: 10.1016/j.physleta.2006.12.041

CrossRef Full Text | Google Scholar

23. Fülöp T. Singular potentials in quantum mechanics and ambiguity in the self-adjoint hamiltonian. SIGMA Symmet Integrabil Geomet Methods Appl. (2007) 3:107. doi: 10.3842/SIGMA.2007.107

CrossRef Full Text

24. Lambert N. String theory 101. In: Baumgartl M, Brunner I, Haack M, editors. Strings and Fundamental Physics. Lecture Notes in Physics. Berlin; Heidelberg: Springer Berlin Heidelberg (2012). p. 1–48. Available online at: https://doi.org/10.1007/978-3-642-25947-0_1

Google Scholar

25. Jackiw R editor. Delta function potentials in two-dimensional and three-dimensional quantum mechanics. In: Diverse Topics in Theoretical and Mathematical Physics. Singapore: World Scientific (1991).

Google Scholar

26. Bethe HA. Theory of disintegration of nuclei by neutrons. Phys Rev. (1935) 47:747–59. doi: 10.1103/PhysRev.47.747

CrossRef Full Text | Google Scholar

27. Landau LD, Lifshitz EM. Chapter XVIII - Inelastic collisions. In: Landau LD, Lifshitz EM, editors. Quantum Mechanics, 3rd Edn. Pergamon (1977). p. 591–646. Available online at: http://www.sciencedirect.com/science/article/pii/B9780080209401500256

28. Audi G, Kondev FG, Wang M, Huang WJ, Naimi S. The NUBASE2016 evaluation of nuclear properties. Chin Phys C. (2017) 41:030001. doi: 10.1088/1674-1137/41/3/030001

CrossRef Full Text | Google Scholar

29. Feldman HA, Kamenshchik AY. Decoherence properties of scalar field perturbations. Class Quant Grav. (1991) 8:L65–71. doi: 10.1088/0264-9381/8/3/003

CrossRef Full Text | Google Scholar

30. Burgess CP, Holman R, Tasinato G, Williams M. EFT beyond the horizon: stochastic inflation and how primordial quantum fluctuations go classical. J High Energy Phys. (2015) 2015:90. doi: 10.1007/JHEP03(2015)090

CrossRef Full Text | Google Scholar

31. Landau LD, Lifshitz EM. Chapter XVII - Elastic Collisions In: Landau LD, Lifshitz EM, editors. Quantum Mechanics, 3rd Edn. Pergamon (1977). p. 502–90.

Google Scholar

Appendix

A. Multi-Particle Partial-Wave Scattering

Here we review the general framework of partial-wave scattering, including discussion of inelastic and multi-channel scattering.

A.1. Single Particle Elastic Scattering

We consider a spinless particle scattering off of a spinless, infinitely massive object at rest at the origin, through a spherically symmetric interaction V(r). As usual, we employ the ansatz that at large distances from the target, the wavefunction of the incident particle is the sum of a plane wave incident along the z-axis and a scattered spherical wave:

ψAns(r)C(eikz+f(θ,ϕ)eikrr).    (A1)

The differential cross-section is the ratio of the flux of the scattered particles Fsc to the flux of the incoming particles Fin. With an incident beam of N particles, the incoming flux is N jin·ez=N|C|2k/m, and the scattered flux is N|C|2|f(θ,ϕ)|2k/mr2, so that and is given by

dσdΩ:=FscFin=1Finjscerr2=| f(θ) |2.    (A2)

And for a spherically symmetric scatterer, f(θ, ϕ) = f(θ).

At the same time, we consider solutions to the full Schrodinger equation

1r2r(r2rψ(r))[ (+1)r2+2mV(r)k2 ]ψ(r)=0,    (A3)

with k2: = 2mE, and the full wavefunction expanded in a series of spherical harmonics Ψ(x,t)=e-iEtψ(r)Y0 (where we set m = 0 due to conservation of angular momentum). The asymptotic form of these radial functions is:

ψSch(r)Aei(krπ/2)2ikr+Bei(krπ/2)2ikr,    (A4)

Finding f(θ) now amounts to matching (A1) and (A4). This can be accomplished by writing the plane-wave eikz in terms of Legendre polynomials. The standard expansion is given as Landau and Lifshitz [31]

eikz==0(2+1)ij(kr)P(cos(θ))       =0(2+1)iei(krπ/2)ei(krπ/2)2ikrP(cos(θ)).    (A5)

Choosing the Condon-Shortly phase convention, the spherical harmonics can be written Y0=(2+1)/4πP(cos(θ)), so by computing the difference:

ψSch(r)Ceikz=Cf(θ)eikrr,    (A6)

one finds first:

B=4π(2+1)iC,    (A7)

set by the fact that there can be no incoming wave in the scattered wavefunction. Finally, one finds

f(θ)=12ik=0(2+1)[ S1 ]P(θ),    (A8)

Where S = −A/B is the scattering matrix element. When the scattering is elastic as we've just described (kout = kin), the matrix element S=e2iδ is a pure phase. Otherwise, when the scattering is inelastic and probability is lost, S is just a complex number, but it is still common [14] to parameterize it as S(in)=e2iγ, where γ is now a complex number.

Finally, the total cross-section is computed as the integral over the differential cross-section, which is (using the orthogonality of the Legendre polynomials)

σ=dΩdσdΩ=πk2(2+1)|S1|2.    (A9)

A.2. Multi-Channel Scattering

It is a simple matter to generalize the above to multi-channel scattering. We treat the case of two species of particles, but as detailed in section 6 the results are more general. Without loss of generality, we will only look at 1 → X scattering.

Following Landau and Lifshitz [27], we begin by assuming the asymptotic forms for each species:

ψ1Ans(r)C(eik1z+f1(θ,ϕ)eik1rr).    (A10)

and

ψ2Ans(r)Cf2(θ,ϕ)eik2rr.    (A11)

The differential cross-sections are defined in exactly the same way as the above. This means the 1 → 1 scattering is exactly given by (A9), while for 1 → 2 scattering we have

dσ12dΩ=k2k1m1m2| f2(θ) |2.    (A12)

Particle 2 satisfies the same Schrödinger equation (A3), and so has the same asymptotic form (A4). Matching to the ansatz is then as simple as

ψ2,Sch=Cf2(θ)eik2rr,    (A13)

which produces

B2=0,    (A14)

and

f2(θ)=12ik2Ceiπ/2A2Y0(θ)            =12ik2(2+1)S12()P(cosθ),    (A15)

which defines the scattering matrix element S12()=e-iπ/2[4π(2+1)]-1/2A2/C. Finally, the total cross-section is

σ12=πk1k2m1m2(2+1)| S12() |2.    (A16)

One important observation: as long as S12()0 for any ℓ, then S11()=e2iγ11() must satisfy that the phase γ11() is complex, since some of the probability flux of the incident particle 1 must have transferred to particle 2.

B. Solving for 2-Particle Integration Constant Ratios

Here we outline the details of the main calculation in section 4.1. We do this for 1 → X scattering, but the results are easily applied to 2 → X scattering by inverting 1↔2.

We begin with the boundary condition (4.2), and the general forms

ψi=Ci+ψi++Ci-ψi-,    (A17)

where i = 1, 2. For 1 → X scattering, we use C2+ = R C2− with R defined in (4.4), and the boundary conditions are:

      ψ^1++C11ψ^1-=h11(ψ1++C11ψ1-)+h12C12                                                   (ψ2+R+ψ2-),and    (A18a)
C12(ψ^2+R+ψ^2-)=h22C12(ψ2+R2+ψ2-)+h21                                                   (ψ1++C11ψ1-).    (A18b)

where we define ψ^i±:=2πϵ2miϵψi±, and as in (4.3), we define C11: = C1−/C1+ and C12: = C2−/C1+.

Rearranging (A18a) for C12, one finds

C12=[ψ^1+-h11ψ1+]+C11[ψ^1--h11ψ1-]h12[ψ2+R+ψ2-].    (A19)

Substituting in (A18b),

{[ψ^1+-h11ψ1+]+C11[ψ^1--h11ψ1-]}Z=|h12|2[ψ1++C11ψ1-],    (A20)

where

Z:=ψ^2+R+ψ^2-ψ2+R+ψ2--h22.    (A21)

Finally rearranging, we have

C11=-ψ1+ψ1-[|h12|2-[ψ^1+ψ1+-h11]Z][|h12|2-[ψ^1-ψ1--h11]Z]    (A22)

Plugging this back into (A19), we have

C12=ψ1+ψ2-[ψ^1+ψ1+-h11][|h12|2-[ψ^1-ψ1--h11]Z]-[+-]h12[Rψ2+ψ2-+1][|h12|2-[ψ^1-ψ1--h11]Z],        =ψ1+ψ2-[ψ^1+ψ1+-ψ^1-ψ1-]h21[Rψ2+ψ2-+1][|h12|2-[ψ^1-ψ1--h11]Z].    (A23)

The integration constants in the 2 → X system are solved for in the same way. Solutions for C22 = C2−/C2+ and C21 = C1−/C2+ in the 2 → X are obtained directly from (A22) and (A23) (respectively) by simply inverting 1↔2.

In order to make use of these formulae, we now have to take the small-r limit of the mode functions to the appropriate order.

B.1. Leading-Order in kiϵ

First, we make the usual leading-order approximation. For ψ1, this is exactly as in (4.6):

      ψ1(ϵ)x1-1/2[C1+x1ζ1/2+C1-x1-ζ1/2],  andrψ1(ϵ)ik1x1-3/2[(ζ1-1)C1+x1ζ1/2-(ζ1+1)C1-x1-ζ1/2],    (A24)

where for convenience here we define xi : = (2ikiϵ), for i = 1, 2. Here we keep both the divergent (−) and the (often) finite (+) term because the ratio C1−/C1+ arises from the point-particle dynamics, and so is of the order kiϵ ≪ 1, which allows the two terms in (A24) to compete. For particle 2 however, this is not the case. C2+ = RC2− with RO(1) so that there is no balancing of the modes, and the divergent mode is simply dominant. That is:

      ψ2(ϵ)C2-x2-1/2-ζ2/2,  andrψ2(ϵ)-(ζ2+1)ik2C2-x2-3/2-ζ2/2.    (A25)

With these approximations, we compute:

ψ^i±ψi±-πϵmi(1ζi),ψ1+ψ1-(2ik1ϵ)ζ1,ψ2+ψ2-0,  and      Zψ^2-ψ2--h22-[(1+ζ2)πϵm2+h22].    (A26)

Then to leading order in kiϵ, it is found that:

C11-(2ik1ϵ)ζ1              [|h12|2-[πϵm1(1-ζ1)+h11][(1+ζ2)πϵm2+h22]][|h12|2-[πϵm1(1+ζ1)+h11][(1+ζ2)πϵm2+h22]],    (A27)

and

C12=(2ik1ϵ)ζ1/2(2ik2ϵ)ζ2/2k2k1            2πϵζ1h21m1[|h12|2-[πϵm1(1+ζ1)+h11][(1+ζ2)πϵm2+h22]]    (A28)

It will become apparent that a redefinition of parameters can significantly clean up our equations. Drawing from the next appendix, we define

λ^11:=1ζ1(m1h11πϵ+1)   λ^12:=h12m1m22πϵζ1ζ2,λ^22:=1ζ2(m2h22πϵ+1)   λ^21:=h21m1m22πϵζ1ζ2.    (A29)

In terms of these new variables, the integration constants read:

C11=-(2ik1ϵ)ζ1[4λ^12λ^21-[λ^11-1] [λ^22+1]][4λ^12λ^21-[λ^11+1] [λ^22+1]],    (A30)

and

C12= (2ik1ϵ)ζ1/2(2ik2ϵ)ζ2/2m2k2ζ1m1k1ζ2             4λ^21[4λ^12λ^21-[λ^11+1] [λ^22+1]].    (A31)

B.2. Sub-Leading-Order in kiϵ

To leading order, C11~(2ikϵ)ζ1, so that to leading order S11() is a pure phase. In order to see the emergence of an absorptive single-particle model in the particle 1 subsector of the theory (as covered in section 7), it is necessary to compute C11 to the next order in kiϵ. To that end, recall:

ψ±=(2ikϵ)12(-1±ζ)e-ikϵM[12(1±ζ),1±ζ,2ikϵ],    (A32)

so that

ψ±(2ikϵ)12(-1±ζ)[1-ikϵ+O((kϵ)2)][1+ikϵ+O((kϵ)2)],       (2ikϵ)12(-1±ζ)[1+O((kϵ)2)]    (A33)

so at least for ζ < 2, the leading correction is only in ψ2, and is to include the (+) mode, since it is only a factor of (k2ϵ)ζ2 compared to the two powers from all other higher-order corrections. In fact this is a property only of systems without a 1/r potential, as in that case the leading correction from the hypergeometric factor does not cancel that from the exponential.

Pushing through then, we repeat the calculation from the previous section, now using

      ψ2(ϵ)C2-x2-1/2[x2-ζ2/2+Rx2ζ2/2],  andrψ2(ϵ)ik2C2-x2-3/2[-(ζ2+1)x2-ζ2/2+(ζ2-1)Rx2ζ2/2].    (A34)

With these approximations, we compute:

ψ^i±ψi±-πϵmi(1ζi),ψi+ψi-(2ikiϵ)ζi,  and       Zψ^2-+Rψ^2+ψ2-+Rψ2+-h22=ψ^2-ψ2-1+Rψ^2+/ψ^2-1+Rψ2+/ψ2--h22         -[(1+ζ2)πϵm2]{1+R(ψ^2+ψ^2--ψ2+ψ2-)}-h22,         =-[(1+ζ2)πϵm2]{1-2ζ2Rζ2+1(2ik2ϵ)ζ2}-h22.    (A35)

Substituting in (A22), one finds

C1--(2ik1ϵ)ζ1{|h12|2-[πϵm1(1-ζ1)+h11][(1+ζ2)πϵm2(1-2ζ2Rζ2+1(2ik2ϵ)ζ2)+h22]}{|h12|2-[πϵm1(1+ζ1)+h11][(1+ζ2)πϵm2(1-2ζ2Rζ2+1(2ik2ϵ)ζ2)+h22]},         =-(2ik1ϵ)ζ1N1+[πϵm1(1-ζ1)+h11]2πϵζ2Rm2(2ik2ϵ)ζ2D+[πϵm1(1+ζ1)+h11]2πϵζ2Rm2(2ik2ϵ)ζ2,         -(2ik1ϵ)ζ1N1D{1+[[πϵm1(1-ζ1)+h11]N1-[πϵm1(1+ζ1)+h11]D]2πϵζ2Rm2(2ik2ϵ)ζ2},    (A36)

where

N1:=|h12|2-[πϵm1(1-ζ1)+h11] [(1+ζ2)πϵm2+h22],  and   D:=|h12|2-[πϵm1(1+ζ1)+h11] [(1+ζ2)πϵm2+h22].    (A37)

Then (A36) simplifies to:

C11=-(2ik1ϵ)ζ1N^1D^{1-4Rλ^12λ^21N1^D^(2ik2ϵ)ζ2}.    (A38)

C. Solving for Point-Particle Couplings

To find the running of the point-particle couplings, we need to isolate for them in the boundary conditions. To do so, we follow the same prescription as B. Write:

ψi=Ci+ψi++Ci-ψi-,    (A39)

where i = 1, 2. We again use (5.8) to define C11 : = C1−/C1+ and C12 : = C2−/C1+ in the 1 → X system, and analogously C22 : = C2−/C2+ and C21 : = C1−/C2+ in the 2 → X system.

For convenience, define the following:

ψ11:=ψ1++C11ψ1-   and   ψ12:=C12[Rψ2++ψ2-]    (A40)

for the 1 → X system, and

ψ22:=ψ2++C22ψ2-   and   ψ21:=C21[Rψ1++ψ1-]    (A41)

for the 2 → X system. The small-r boundary conditions (4.5) and (4.10) can be written

ψ^11=h11ψ11+h12ψ12,  and    (A42)
ψ^12=h22ψ12+h21ψ11,    (A43)

and

ψ^21=h11ψ21+h12ψ22,  and    (A44)
ψ^22=h22ψ22+h21ψ21,    (A45)

As in Appendix B, we define ψ^:=4πϵ2ϵψ. Using (A42) and (A44), we can isolate for h11 and h12:

h11=ψ^11ψ22-ψ^21ψ12ψ11ψ22-ψ21ψ12   and   h12=ψ^11ψ21-ψ^21ψ11ψ12ψ21-ψ22ψ11.    (A46)

Similarly, using (A43) and (A45), we can isolate for h22 and h21:

h22=ψ^12ψ21-ψ^22ψ11ψ12ψ21-ψ22ψ11   and   h21=ψ^12ψ22-ψ^22ψ12ψ11ψ22-ψ21ψ12.    (A47)

To make use of these formulae, we approximate ψij using the small-r forms as used in Appendix B:

                      ψ11(22)(ϵ)x1(2)-1/2[x1(2)ζ1(2)/2+C11(22)x1(2)-ζ1(2)/2],  and4πϵ2r|ϵψ11(22)(ϵ)2πϵx1(2)-1/2[(ζ1(2)-1)x1(2)ζ1(2)/2-(ζ1(2)+1)C11(22)x1(2)-ζ1(2)/2],    (A48)

and

                     ψ21(12)(ϵ)C21(12)x1(2)-1/2-ζ1(2)/2,  and4πϵ2r|ϵψ21(12)(ϵ)-(ζ1(2)+1)2πϵC21(12)x1(2)-1/2-ζ1(2)/2.    (A49)

where again, xi : = (2ikiϵ), with i = 1, 2. Substituting (A48) and (A49) into (A46) and (A47), we have the following. For h11 we find

λ^11=(x1ζ1/2-C11x1-ζ2/2)(x2ζ1/2+C22x2-ζ2/2)+C21C12x1-ζ1/2x2-ζ2/2(x1ζ1/2+C11x1-ζ2/2)(x2ζ1/2+C22x2-ζ2/2)-C21C12x1-ζ1/2x2-ζ2/2,          =(1-C11x1-ζ1)(1+C22x2-ζ2)+C21C12x1-ζ1x2-ζ2(1+C11x1-ζ1)(1+C22x2-ζ2)-C21C12x1-ζ1x2-ζ2,    (A50)

which defines λ^11:=1ζ1(m1h11πϵ+1). Notice the limit C21 = C12 → 0 reduces λ^11 to the single-species running (2.9), as it should (the limit in which there is no mixing between particle species). For h12, we have

λ^12=m2ζ1m1ζ2C21(x1x2)-1/2(x1ζ1/2+C11x1-ζ2/2)(x2ζ1/2+C22x2-ζ2/2)-C21C12x1-ζ1/2x2-ζ2/2,          =m2ζ1m1ζ2C21(x1x2)-1/2x1-ζ1/2x2-ζ2/2(1+C11x1-ζ1)(1+C22x2-ζ2)-C21C12x1-ζ1x2-ζ2,    (A51)

with now λ^12:=h12m1m22πϵζ1ζ2. Notice again the clean limit h12 → 0 when C21 → 0. The rest follow easily:

λ^21=m1ζ2m2ζ1C12(x2x1)-1/2x1-ζ1/2x2-ζ2/2(1+C11x1-ζ1)(1+C22x2-ζ2)-C21C12x1-ζ1x2-ζ2,    (A52)

similarly with λ^21:=h21m1m22πϵζ1ζ2. Lastly,

λ^22=(1+C11x1-ζ1)(1-C22x2-ζ2)+C21C12x1-ζ1x2-ζ2(1+C11x1-ζ1)(1+C22x2-ζ2)-C21C12x1-ζ1x2-ζ2,    (A53)

with λ^22:=1ζ2(m2h222πϵ+1).

Keywords: flavor violation, effective field theories, catalysis, high energy physics - theory, nuclear theory

Citation: Hayman P and Burgess CP (2019) Point-Particle Catalysis. Front. Phys. 7:167. doi: 10.3389/fphy.2019.00167

Received: 30 April 2019; Accepted: 11 October 2019;
Published: 05 November 2019.

Edited by:

Manuel Gadella, University of Valladolid, Spain

Reviewed by:

Andreas Gustavsson, University of Seoul, South Korea
Manu Paranjape, Université de Montréal, Canada

Copyright © 2019 Hayman and Burgess. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Peter Hayman, haymanpf@mcmaster.ca

Download