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

Nucleon electric dipole form factor in QCD vacuum

Wei-Yang Liu [email protected] Center for Nuclear Theory, Department of Physics and Astronomy, Stony Brook University, Stony Brook, New York 11794-3800, USA    Ismail Zahed [email protected] Center for Nuclear Theory, Department of Physics and Astronomy, Stony Brook University, Stony Brook, New York 11794-3800, USA
Abstract

In the QCD instanton vacuum, the C-even and C-odd nucleon Pauli form factors receive a large contribution from the underlying ensemble of pseudoparticles, that is sensitive to a finite QCD vacuum angle θ\theta. This observation is used to derive explicitly the electric dipole form factor for light quark flavors, and estimate the proton and neutron electric dipole moment induced by a small CP violating θ\theta. The results are in the range of some recently reported lattice simulations.

I Introduction

The resolution of the CP problem in baryogenesis is central to our understanding of baryon asymmetry in the universe Sakharov (1967). CP violation in the weak sector of the standard model, proves to be orders of magnitude away from its resolution, while CP violation in the strong sector, puts its resolution within range Kharzeev et al. (2020).

The QCD vacuum is rich with topologically active pseudoparticles which are CP conjugated pairs (instantons and anti-instantons). They are natural sources of local CP violation effects. CP is strongly violated in QCD at finite theta angle. A reliable description of ensembles of these pseudoparticles in the semi-classical approximation, is provided by the instanton liquid model (ILM) Diakonov et al. (1996); Schäfer and Shuryak (1998); Nowak et al. (1996); Liu (2025) (and references therein). The model has proven to be very useful in capturing many aspects of most hadronic correlations both in vacuum, in hadronic states and in matter. Here, we will focus on understanding the role of these pseudoparticles in the composition of the hadronic electric dipole moment.

For many decades, the nucleon electric dipole moment has been used as a measure of the strong CP violation caused by a finite theta angle in QCD. The current empirical estimate puts its upper bound at about 1026e10^{-26}\,e-cm Abel et al. (2020). This is an ideal task for ab-initio lattice simulations, yet the smallness of the observable makes the task daunting in light of the signal-to-noise ratio Syritsyn et al. (2019); Alexandrou et al. (2021). Notwithstanding this, the latest lattice estimate puts it at about 1016θe10^{-16}\,\theta\,e-cm for a finite theta angle Liang et al. (2023), which would limit θ\theta to about 101010^{-10} by the empirical bound.

The QCD vacuum breaks conformal symmetry, a mechanism at the origin of most hadronic masses. Detailed gradient flow (cooling) techniques have revealed a striking semi-classical landscape made of instantons and anti-instantons, the vacuum tunneling pseudoparticles with unit topological charges Leinweber (1999); Michael and Spencer (1995a, b); Biddle et al. (2018); Athenodorou et al. (2018); Ringwald and Schrempp (1999). These pseudoparticles break chiral symmetry through fermionic zero modes with fixed chirality (left or right). The key features of this landscape are Shuryak (1982)

nI+A1R41fm4ρR13n_{I+A}\equiv\frac{1}{R^{4}}\approx\frac{1}{{\rm fm}^{4}}\qquad\qquad\frac{\rho}{R}\approx\frac{1}{3} (1)

for the instanton plus anti-instanton density and size, respectively. The hadronic scale R=1fmR=1\,{\rm fm} emerges as the mean quantum tunneling rate of the pseudoparticles. Deep in the cooling time the tunnelings are sparse, well described by the instanton liquid model (ILM) with a packing fraction

κπ2ρ4nI+A0.1\displaystyle\kappa\equiv\pi^{2}\rho^{4}n_{I+A}\approx 0.1 (2)

The main purpose of this work is to evaluate the C-odd nucleon electric dipole form factor, and provide an estimate of the proton and neutron electric dipole moments in the ILM. In section II we briefly review the salient features of the pseudoparticle fluctuations in the ILM, with a comparison to some lattice simulations. In section III we define the four C-even or odd form factors associated to the electric current in the nucleon. The C-odd electric dipole form factor for each flavor contribution is evaluated in the ILM. This form factor is found to be related to the Pauli form factor. In particular, the finite proton and neutron electric dipole moments in the ILM are shown to be comparable to some recent lattice simulations. Our conclusions are in section IV. Further details can be found in the appendices.

Refer to caption
Figure 1: The ILM results unquenched (red-solid) and quenched (blue-solid), compared to the lattice results from the ETMC collaboration Alexandrou et al. (2021).

II QCD instanton vacuum

The size distribution of the instantons and anti-instantons density (their tunneling rate) in the vacuum is well captured semi-empirically by

n(ρ)1ρ5(ρΛQCD)beCρ2/R2n(\rho)\sim{1\over\rho^{5}}\big{(}\rho\Lambda_{QCD}\big{)}^{b}\,e^{-C\rho^{2}/R^{2}} (3)

with the mean value (1). Here b=11Nc/32Nf/3b=11N_{c}/3-2N_{f}/3 (one loop), and CC is a number of order 1 fixed by the binary gauge interactions among pseudoparticles in the vacuum Diakonov (1996); Shuryak (1999).

To capture the fluctuations of the number of pseudoparticles N±N_{\pm}, we identify them with the scalar and pseudoscalar gluonic densities in the vacuum

132π2d4xF2(x)\displaystyle\frac{1}{32\pi^{2}}\int d^{4}x\,F^{2}(x) =\displaystyle= (N++N)=N\displaystyle(N_{+}+N_{-})=N
132π2d4xFF~(x)\displaystyle\frac{1}{32\pi^{2}}\int d^{4}x\,F\tilde{F}(x) =\displaystyle= (N+N)=Δ\displaystyle(N_{+}-N_{-})=\Delta (4)

with gFFgF\rightarrow F assumed. For a random quenched vacuum with Poissonian N±N_{\pm} at finite angle θ\theta,

F2θ32π2\displaystyle\frac{\langle F^{2}\rangle_{\theta}}{32\pi^{2}} \displaystyle\approx nI+Acosθ\displaystyle n_{I+A}\,{\rm cos\,\theta}
FF~θ32π2\displaystyle\frac{\langle F\tilde{F}\rangle_{\theta}}{32\pi^{2}} \displaystyle\approx inI+Asinθ\displaystyle in_{I+A}\,{\rm sin\,\theta} (5)

with the breaking of scale or conformal symmetry manifested for vanishing theta. However, in the ILM the distributions of N±N_{\pm} are not Poissonian. The fluctuations in the sum NN follow from low-energy theorems Novikov et al. (1981)

(N)=ebN¯4(N¯N)bN4\mathbb{P}(N)=e^{-\frac{b\bar{N}}{4}}\bigg{(}\frac{\overline{N}}{N}\bigg{)}^{\frac{bN}{4}} (6)

with the vacuum topological compressibility

σTV4=(NN¯)2V4=4bnI+A\frac{\sigma_{T}}{V_{4}}=\frac{\langle(N-\bar{N})^{2}\rangle}{V_{4}}=\frac{4}{b}\,n_{I+A} (7)

and the volume extensive mean N¯=nI+AV4\bar{N}=n_{I+A}V_{4}. The fluctuations in the difference Δ\Delta are fixed by the topological susceptibity Diakonov et al. (1996). At θ=0\theta=0, the distribution reads

(Δ)=12πχtexp(Δ22χt)\mathds{Q}(\Delta)=\frac{1}{\sqrt{2\pi\chi_{t}}}\exp\left(-\frac{\Delta^{2}}{2\chi_{t}}\right) (8)

The topological susceptibility χt\chi_{t} in gluodynamics is

χtV4=Δ2V4nI+A\displaystyle\frac{\chi_{t}}{V_{4}}=\frac{\langle\Delta^{2}\rangle}{V_{4}}\approx n_{I+A} (9)

which is large. However, in QCD it is substantially screened by the light quarks, as illustrated in Fig. 1 with current quark mass m=10m=10 MeV. The width of the distribution reads

Δ2N¯(1+Nfmm)1\displaystyle\frac{\langle\Delta^{2}\rangle}{\bar{N}}\sim\bigg{(}1+N_{f}\frac{m^{*}}{m}\bigg{)}^{-1} (10)

Note that the topological susceptibility is sensitive to the determinantal mass mm^{*}. The details about how to fix the determinantal mass are given in Appendix C and in Liu et al. (2023a, 2024) (and references therein).

In Fig. 1 we show the unquenched ILM results (10) (red-solid) and quenched ILM (9) (blue-solid), compared to the lattice results from the ETMC collaboration Alexandrou et al. (2021). The latters used twisted mass clover-improved fermions Nf=2+1+1N_{f}=2+1+1, in a 4-volume 643×128a464^{3}\times 128~{}a^{4} with lattice spacing a=0.0801(4)a=0.0801(4) fm, and physical pion mass mπ=139m_{\pi}=139 MeV. For the ILM parameters see below.

III Nucleon EM form factors

At finite vacuum angle θ\theta, the QCD action in Euclidean signature is supplemented by the topological term

θ32π2d4xFF~θΔ\frac{\theta}{32\pi^{2}}\int d^{4}x\,F\tilde{F}\rightarrow\theta\Delta

which is at the origin of strong CP violation. In the ILM, this contribution acts as a topological chemical potential for Δ\Delta, by enhancing instantons and depleting anti-instantons. This affects most observables, and in particular the electromagnetic (EM) form factors of hadrons, as we will detail in this section.

The coupling to photon can be described by Dirac F1F_{1}, Pauli F2F_{2}, electric dipole moment form factor F3F_{3}, and axial tensor form factor FAF_{A}.

N|JEMμ=fQfψ¯fγμψf|N\displaystyle\langle N^{\prime}|J^{\mu}_{\rm EM}=\sum_{f}Q_{f}\bar{\psi}_{f}\gamma^{\mu}\psi_{f}|N\rangle (11)
=u¯s(P)[γμF1(Q2)+iσμνqν2MN(F2(Q2)iγ5F3(Q2))+1MN2(qμq2γμ)γ5FA(Q2)]us(P)\displaystyle=\bar{u}_{s^{\prime}}(P^{\prime})\left[\gamma^{\mu}F_{1}(Q^{2})+\frac{i\sigma^{\mu\nu}q_{\nu}}{2M_{N}}\left(F_{2}(Q^{2})-i\gamma^{5}F_{3}(Q^{2})\right)+\frac{1}{M_{N}^{2}}\left(\not{q}q^{\mu}-q^{2}\gamma^{\mu}\right)\gamma^{5}F_{A}(Q^{2})\right]u_{s}(P)

with the Sachs FFs GC,MG_{C,M}

GC(Q2)\displaystyle G_{C}(Q^{2}) =\displaystyle= F1(Q2)+Q24MN2F2(Q2)\displaystyle F_{1}(Q^{2})+\frac{Q^{2}}{4M^{2}_{N}}F_{2}(Q^{2})
GM(Q2)\displaystyle G_{M}(Q^{2}) =\displaystyle= F1(Q2)+F2(Q2)\displaystyle F_{1}(Q^{2})+F_{2}(Q^{2})\, (12)

and the electric dipole and axial-tensor FFs GD,AG_{D,A}

GD(Q2)\displaystyle G_{D}(Q^{2}) =\displaystyle= F3(Q2)\displaystyle F_{3}(Q^{2})
GA(Q2)\displaystyle G_{A}(Q^{2}) =\displaystyle= FA(Q2)\displaystyle F_{A}(Q^{2}) (13)

with the two latters vanishing for θ=0\theta=0. The magnetic and electric dipole moment are defined respectively,

μN=GM(0)e/2MN\displaystyle\mu_{N}=G_{M}(0)e/2M_{N} dN=F3(0)e/2MN\displaystyle d_{N}=F_{3}(0)e/2M_{N} (14)

For finite theta, (LABEL:EM_form) will be organized through

JμDirac(q,θ)+JμPauli(q,θ)+JμA(q,θ)J_{\mu}^{\mathrm{Dirac}}(q,\theta)+J_{\mu}^{\mathrm{Pauli}}(q,\theta)+J_{\mu}^{A}(q,\theta) (15)
Refer to caption
Figure 2: Leading pseudoparticle (single instanton) contribution to the quark EM current in the ILM.

III.1 Emergent EM vertex

To see how the pseudoparticles contribute to the EM FF of hadrons, consider the instanton (antiinstanton) insertion to the charge current on a quark line illustrated in Fig. 2. In an instanton (antiinstanton) the incoming left-handed (right-handed) quark flips to a right-handed (left-handed) through a zero mode, then scatters off a virtual photon before exiting through a non-zero mode, with the result for a single quark Kochelev (2003). The effective interaction amplitude is defined as

d4zeiqz\displaystyle\int d^{4}ze^{-iq\cdot z} [SNZM(x,z)γμSZM(z,y)\displaystyle\bigg{[}S_{\mathrm{NZM}}(x,z)\gamma^{\mu}S_{\mathrm{ZM}}(z,y) (16)
+SZM(x,z)γμSNZM(z,y)]\displaystyle+S_{\mathrm{ZM}}(x,z)\gamma^{\mu}S_{\mathrm{NZM}}(z,y)\bigg{]}

For a single instanton, the zero mode propagator is

SZM(x,x)=ϕI(x)ϕI(x)imϕI(x)ϕI(x)imS_{\mathrm{ZM}}(x,x^{\prime})=\frac{\phi_{I}(x)\phi_{I}^{\dagger}(x^{\prime})}{im}\rightarrow\frac{\phi_{I}(x)\phi_{I}^{\dagger}(x^{\prime})}{im^{*}} (17)

with the zero modes defined in (37). The singular 1/m1/m in the single instanton approximation is shifted to finite 1/m1/m^{*} by disordering in the ILM Pobylitsa (1989); Schäfer and Shuryak (1998); Liu (2025) (and references therein). The determinantal mass mm^{*} is seen to follow from the constituent mass through a similar disordering Liu et al. (2024)

Mq(0)=nI+A2Nc4π2ρ2mnI+A2Nc4π2ρ2mM_{q}(0)=\frac{n_{I+A}}{2N_{c}}\frac{4\pi^{2}\rho^{2}}{m}\rightarrow\frac{n_{I+A}}{2N_{c}}\frac{4\pi^{2}\rho^{2}}{m^{*}} (18)

with the identification

m=\displaystyle m^{*}= nI+A2Nc(2qφ2)\displaystyle\sqrt{\frac{n_{I+A}}{2N_{c}}}\left(\sqrt{2}||q\varphi^{\prime 2}||\right)

The norm uses the zero mode profile φ(q)\varphi^{\prime}(q) defined as

φ(q)=πρ2(I0K0(z)I1K1(z))z=ρq/2\varphi^{\prime}(q)=\pi\rho^{2}\bigg{(}I_{0}K_{0}(z)-I_{1}K_{1}(z)\bigg{)}^{\prime}_{z=\rho q/2} (20)

The typical value of the constituent mass is about Mq(0)380395M_{q}(0)\approx 380-395 MeV Liu et al. (2023a, b); Liu (2025).

Inserting (17) into (16) yield the instanton induced effective EM vertex

V+μ(x,y)=d4zI𝑑UId4zeiqz\displaystyle V^{\mu}_{+}(x,y)=\int d^{4}z_{I}dU_{I}\int d^{4}ze^{-iq\cdot z} [S¯nz(I)(xzI,zzI)γμϕI(zzI)ϕI(yzI)im\displaystyle\bigg{[}\bar{S}^{(I)}_{nz}(x-z_{I},z-z_{I})\gamma^{\mu}\frac{\phi_{I}(z-z_{I})\phi_{I}^{\dagger}(y-z_{I})}{im^{*}} (21)
+ϕI(xzI)ϕI(zzI)imγμSnz(I)(zzI,yzI)]1+γ52\displaystyle+\frac{\phi_{I}(x-z_{I})\phi_{I}^{\dagger}(z-z_{I})}{im^{*}}\gamma^{\mu}S^{(I)}_{nz}(z-z_{I},y-z_{I})\bigg{]}\frac{1+\gamma^{5}}{2}

The effective vertex for the anti-instanton Vμ(x,y)V^{\mu}_{-}(x,y) follows by interchanging τμτμ+\tau^{-}_{\mu}\leftrightarrow\tau^{+}_{\mu}, σμσ¯μ\sigma_{\mu}\leftrightarrow\bar{\sigma}_{\mu}, and γ5γ5\gamma^{5}\leftrightarrow-\gamma^{5}. Now the effective quark operator can be written as

d4k(2π)4d4k(2π)4ψ¯(k)[N+VV+(k,k)+NVV(k,k)]ψ(k)\int\frac{d^{4}k}{(2\pi)^{4}}\frac{d^{4}k^{\prime}}{(2\pi)^{4}}\bar{\psi}(k^{\prime})\left[\frac{N_{+}}{V}V_{+}(k^{\prime},k)+\frac{N_{-}}{V}V_{-}(k^{\prime},k)\right]\psi(k) (22)

The on-shell reduction of the in-out quark lines can be obtained by using the large time asymptotics as detailed in appendix B.2. More specifically, the on-shell reduction of the emergent vertex is (q2=Q2<0q^{2}=-Q^{2}<0)

V+(k,k)=limx4+y4[d3xd3yeikxiγ4V+μ(x,y)iγ4eiky]ik4=|k|ik4=|k|\displaystyle V_{+}(k^{\prime},k)=\lim_{\begin{subarray}{c}x_{4}\rightarrow+\infty\\ y_{4}\rightarrow-\infty\end{subarray}}\left[\int d^{3}\vec{x}d^{3}\vec{y}e^{ik^{\prime}\cdot x}i\gamma_{4}V^{\mu}_{+}(x,y)i\gamma_{4}e^{-ik\cdot y}\right]_{\begin{subarray}{c}ik_{4}=|\vec{k}|\\ ik^{\prime}_{4}=|\vec{k}^{\prime}|\end{subarray}} (23)

with the result for the Pauli contribution

V±(k,k)\displaystyle V_{\pm}(k^{\prime},k) \displaystyle\simeq i(2π)4δ(kk+q)8π2ρ4Nc1γ52iσμνqν2m\displaystyle~{}i(2\pi)^{4}\delta(k-k^{\prime}+q)\frac{8\pi^{2}\rho^{4}}{N_{c}}\frac{1\mp\gamma^{5}}{2}\frac{i\sigma^{\mu\nu}q_{\nu}}{2m^{*}} (24)
×01dt[tK0(u1t)181tuu(uK1(u1t))]|u=ρQ\displaystyle\times\int_{0}^{1}dt\left[tK_{0}(u\sqrt{1-t})-\frac{1}{8}\frac{\sqrt{1-t}}{u}\frac{\partial}{\partial u}\left(uK_{1}(u\sqrt{1-t})\right)\right]\bigg{|}_{u=\rho Q}

where Kn(x)K_{n}(x) is the modified Bessel function of the second kind.

III.2 ff-quark EDM

At zero vacuum angle, the Pauli contribution to the charge form factor in (15), receives a large contribution from pseudoparticles in the QCD instanton vacuum Kochelev (2003). We now show that this observation carries to the CP-odd contributions at finite vacuum angle. The Pauli part of the form factor of a quark of flavor ff is then

Jμ,fPauli(q,θ)\displaystyle J^{\rm Pauli}_{\mu,f}(q,\theta) =\displaystyle= QfeFI(ρQ)u¯s(k)(N+N¯1γ52iσμνqν2m+NN¯1+γ52iσμνqν2m)us(k)\displaystyle Q_{f}eF_{I}(\rho Q)\bar{u}_{s}(k^{\prime})\bigg{(}\frac{N_{+}}{\bar{N}}\frac{1-\gamma^{5}}{2}\frac{i\sigma_{\mu\nu}q_{\nu}}{2m^{*}}+\frac{N_{-}}{\bar{N}}\frac{1+\gamma^{5}}{2}\frac{i\sigma_{\mu\nu}q_{\nu}}{2m^{*}}\bigg{)}u_{s}(k) (25)

with the pseudoparticle induced form factor

FI(q)=8π2ρ4nI+ANc1q2[(22qK1(q))(7q274qK1(q)72K2(q))]\displaystyle F_{I}(q)=\frac{8\pi^{2}\rho^{4}n_{I+A}}{N_{c}}\frac{1}{q^{2}}\left[\bigg{(}2-2qK_{1}(q)\bigg{)}-\left(\frac{7}{q^{2}}-\frac{7}{4}qK_{1}(q)-\frac{7}{2}K_{2}(q)\right)\right] (26)

The first parenthesis is identical to the Kochelev’s computation using an on-shell limit Kochelev (2003). The additional contribution in our case stems from use of the large Euclidean time approximation in the on-shell reduction scheme discussed in Appendix B. The latter is more appropriate for Euclidean formulations Liu and Zahed (2021).

The averaging of (25) over N±N_{\pm} is carried using the grand-canonical ensemble detailed in appendix D, with the result

Jμ,fPauli(q,θ)=QfeFI(ρQ)u¯s(k)(1Δ2cN¯θiγ5)iσμνqν2mus(k)\displaystyle\langle J^{\rm Pauli}_{\mu,f}(q,\theta)\rangle=Q_{f}eF_{I}(\rho Q)\bar{u}_{s}(k^{\prime})\left(1-\frac{\langle\Delta^{2}\rangle_{c}}{\bar{N}}\theta\,i\gamma^{5}\right)\frac{i\sigma_{\mu\nu}q_{\nu}}{2m^{*}}u_{s}(k) (27)

where we have used N±=N¯±iΔ2cθ\langle N_{\pm}\rangle=\bar{N}\pm i\langle\Delta^{2}\rangle_{c}\theta

It approaches to the asymptotic form in the large momentum transfer limit qq\rightarrow\infty. Here Δ2c=χt\langle\Delta^{2}\rangle_{c}=\chi_{t} is the vacuum connected topological susceptibility (10). A comparison of (26) to (LABEL:EM_form) yields the pseudoparticle contribution to the F2,3F_{2,3} ff-quark form factors

F2f(Q2)\displaystyle F^{f}_{2}(Q^{2}) \displaystyle\rightarrow QfeFI(ρQ)\displaystyle Q_{f}eF_{I}(\rho Q)
F3f(Q2)\displaystyle F^{f}_{3}(Q^{2}) \displaystyle\rightarrow QfeFI(ρQ)Δ2N¯θ\displaystyle Q_{f}eF_{I}(\rho Q)\frac{\langle\Delta^{2}\rangle}{\bar{N}}\theta (28)

in leading order in the density. Note that our result for the CP even F2fF_{2}^{f} form factor is different from the one in Kochelev (2003) which is IR singular. The reason is that our use of the large time asymptotic method presented in appendix B.2 is more appropriate for the on-shell reduction in Euclidean space as emphasized in Liu and Zahed (2021). The logarithmic sensitivity of the CP even F2fF_{2}^{f} form factor, will be fixed by the magnetic moment (see below).

Inserting (26) in (III) yields the ff-quark electric dipole moment

F3f(Q2)θ=\displaystyle\frac{F^{f}_{3}(Q^{2})}{\theta}= (1+Nfmm)1F2f(Q2)\displaystyle\bigg{(}1+N_{f}\frac{m^{*}}{m}\bigg{)}^{-1}F^{f}_{2}(Q^{2}) (29)

for small vacuum angle. The numerical result is sensitive to the value of the emergent determinantal mass mm^{*} given in (III.1).

Table 1: proton and neutron EDM.
Neutron (103eθ10^{-3}~{}e\thetafm\cdot\rm fm) Proton (103eθ10^{-3}~{}e\thetafm\cdot\rm fm) Ratio |dn/dp||d_{n}/d_{p}|
ILM dn=4.73d_{n}=-4.73 dp=4.44d_{p}=4.44 1.0671.067
Faccioli et al. Faccioli et al. (2004) |dn|=614|d_{n}|=6\sim 14 - -
ChPT Mereghetti et al. (2011) |dn|=2.10|d_{n}|=2.10 |dp|=2.38|d_{p}|=2.38 0.8820.882
χ\chiQCDLiang et al. (2023) dn=1.480.31+0.14d_{n}=-1.48^{+0.14}_{-0.31} dp=3.80.8+1.1d_{p}=3.8^{+1.1}_{-0.8} 0.390.12+0.120.39^{+0.12}_{-0.12}
Bhattacharya et al.Bhattacharya et al. (2021) dn=320+7d_{n}=-3^{+7}_{-20} dp=2430+10d_{p}=24^{+10}_{-30} 0.130.30+0.800.13^{+0.80}_{-0.30}
Dragos et al. Dragos et al. (2021) dn=1.52±0.71d_{n}=-1.52\pm 0.71 dp=1.1±1.0d_{p}=1.1\pm 1.0 1.4±1.41.4\pm 1.4
ETMCAlexandrou et al. (2021) |dn|=0.9±2.4|d_{n}|=0.9\pm 2.4 - -

III.3 Proton and neutron EDM

In the ILM, the proton and neutron are strongly correlated quark-diquark states, with a tight scalar-iso-scalar diquark [ud]S[ud]_{S} and weaker axial-vector flavor-triplet diquark [ud]A[ud]_{A} Schäfer and Shuryak (1998). This strong correlation in the scalar channel follows from the particle-anti-particle symmetry between the spin-0 pion and the spin-0 diquark. To take advantage of this observation, the proton and neutron SU(6) wavefunctions can be repacked in quark-diquark contributions Anselmino et al. (1993) (and references therein)

p\displaystyle p\uparrow =\displaystyle= 118(3[ud]Su+2[uu]A+d2[uu]A0d2[ud]A+u+[ud]A0u)\displaystyle\frac{1}{\sqrt{18}}\bigg{(}3[ud]_{S}u\uparrow+2[uu]^{+}_{A}d\downarrow-\sqrt{2}[uu]^{0}_{A}d\uparrow-\sqrt{2}[ud]^{+}_{A}u\downarrow+[ud]^{0}_{A}u\uparrow\bigg{)}
n\displaystyle n\uparrow =\displaystyle= 118(3[ud]Sd+2[dd]A+u2[dd]A0u2[ud]A+d+[ud]A0d)\displaystyle\frac{1}{\sqrt{18}}\bigg{(}3[ud]_{S}d\uparrow+2[dd]^{+}_{A}u\downarrow-\sqrt{2}[dd]^{0}_{A}u\uparrow-\sqrt{2}[ud]^{+}_{A}d\downarrow+[ud]^{0}_{A}d\uparrow\bigg{)} (30)

with the upper labels 0,±0,\pm referring to the vector helicities. We can use (III.3) to evaluate the individual flavor contributions to the CP-odd contribution to the Pauli form factor. For that we note in the ILM the non-zero mode insertion can only happen on the unpaired quark, as the paired quarks in a diquark are locked by zero modes only.

Alternatively, we may use (29) to derive the CP-odd contribution in a proton and neutron through the Pauli form factor, hence

dp(0)θ\displaystyle\frac{d_{p}(0)}{\theta} =\displaystyle= (1+Nfmm)1×(μpQpe2MN)\displaystyle\bigg{(}1+N_{f}\frac{m^{*}}{m}\bigg{)}^{-1}\times\left(\mu_{p}-Q_{p}\frac{e}{2M_{N}}\right)
dn(0)θ\displaystyle\frac{d_{n}(0)}{\theta} =\displaystyle= (1+Nfmm)1×(μnQne2MN)\displaystyle\bigg{(}1+N_{f}\frac{m^{*}}{m}\bigg{)}^{-1}\times\left(\mu_{n}-Q_{n}\frac{e}{2M_{N}}\right) (31)

(27), (29) and (III.3) are the main results of this work.

Using the empirical values μp=2.793e/2MN0.2937\mu_{p}=2.793\ e/2M_{N}\simeq 0.2937 e\cdotfm and μn=1.913e/2MN0.2012\mu_{n}=-1.913\ e/2M_{N}\simeq-0.2012 e\cdotfm Patrignani et al. (2016), the electric dipole moments read

dp(0)eθ\displaystyle\frac{d_{p}(0)}{e\theta} =\displaystyle= +0.004436fm\displaystyle+0.004436\,\rm fm
dn(0)eθ\displaystyle\frac{d_{n}(0)}{e\theta} =\displaystyle= 0.004732fm\displaystyle-0.004732\,\rm fm (32)

with a fixed ratio dp(0)/dn(0)=0.9373d_{p}(0)/d_{n}(0)=-0.9373. The comparison to the recently reported lattice results and chiral perturbation calculation (ChPT) for the nucleon EDM are given in Table 1, with Nf=2N_{f}=2, m=8.0MeVm=8.0~{}\mathrm{MeV} and m=110.7MeVm^{*}=110.7~{}\mathrm{MeV}. Overall, our results appear to be in the range of some of the proton and neutron EDM reported by some lattice collaborations. For completeness, we note an earlier estimate of the neutron dipole moment |dn(0)|=(614)×103(eθfm)|d_{n}(0)|=(6-14)\times 10^{-3}\,(e\theta\cdot\rm fm) using a numerical ensemble of pseudo-particles to describe the ILM, with a moment approximation to extract the EDM Faccioli et al. (2004).

In Fig. 3 we show our result for the C-odd neutron electric dipole form factor (solid-red band) at low Q2Q^{2}. The logarithmic infrared sensitivity of the induced pseudo-particle form factor in (26), is cutoff around 11 fm-1 0.1973\simeq 0.1973 GeV (constant red dashed line) as we start to probe multi-pseudoparticle correlations, well beyond our current single instanton approach in the dilute ILM description. The momentum dependence is compared to the reported lattice C-odd electric dipole form factor (blue data) from the χ\chiQCD lattice collaboration Liang et al. (2023). The green band is their estimated extrapolation using linear fits, and the yellow band is their estimated extrapolation using a square fit with additional Q4Q^{4} term. The lattice simulations are carried out using overlap fermions, in a 4-volume 64×24364\times 24^{3} ensemble with a lattice spacing a=0.1105fma=0.1105\,\rm fm.

Refer to caption
Figure 3: The C-odd electric dipole form factor from the ILM (solid-red) with F3(0)F_{3}(0) value normalized to the lattice linear and square fit. The Q2Q^{2} dependence is compared to the lattice results (blue data) from the χ\chiQCD collaboration Liang et al. (2023) with mπ=339m_{\pi}=339 MeV. The green band is a lattice linear fit and the yellow band is a lattice square fit with additional Q4Q^{4} term

IV Conclusions

In QCD, the breaking of conformal symmetry puts stringent constraints on the bulk hadronic correlations in the form of low energy theorems Novikov et al. (1981). These constraints are enforced in the QCD instanton vacuum in the form of stronger than Poisson fluctuations in the number of pseudoparticles, with a vacuum compressibility indicative of a quantum liquid. The fluctuations in the difference of the pseudoparticles are peaked around neutral topological charge with the variance, which is fixed by the topological susceptibility which is large in gluodynamics, but substantially screened in QCD. These are essential features of the ILM.

In the ILM, the Pauli form factor of constituent quarks receive a large contribution in the ILM, correcting an early observation in Kochelev (2003). In the presence of a CP violation by a small vacuum angle θ\theta, we have shown that an axial contribution develops in the Pauli form factor, driven mostly by the vacuum topological susceptibility. We have used this result to estimate the induced the electric dipole moment in both the proton and neutron, as well as the semi-hard momentum dependence of the Pauli form factors F2F_{2}, F3F_{3} induced by the pseudoparticles. Our results are in the range of recently reported lattice results extrapolated at the physical pion mass Liang et al. (2023).


Acknowledgements

We thank Fangcheng He for some discussions. This work is supported by the Office of Science, U.S. Department of Energy under Contract No. DE-FG-88ER40388. This research is also supported in part within the framework of the Quark-Gluon Tomography (QGT) Topical Collaboration, under contract no. DE-SC0023646.

Appendix A Quark propagator in the single instanton background

In singular gauge, the instanton gauge field is given by

Aμa(x;ΩI)=1gRab(UI)η¯μνbνlnΠ(xzI)A^{a}_{\mu}(x;\Omega_{I})=-\frac{1}{g}R^{ab}(U_{I})\bar{\eta}^{b}_{\mu\nu}\partial_{\nu}\ln\Pi(x-z_{I}) (33)

Here the instanton moduli is captured by ΩI=(zI,ρ,UI)\Omega_{I}=(z_{I},\rho,U_{I}) the rigid color rotation UIU_{I}, instanton location zIz_{I} and size ρ\rho, with the singular gauge potential

Π(x)=1+ρ2x2\Pi(x)=1+\frac{\rho^{2}}{x^{2}} (34)

The rigid color rotation

Rab(UI)=12Tr(τaUIτbUI)R^{ab}(U_{I})=\frac{1}{2}\mathrm{Tr}(\tau^{a}U_{I}\tau^{b}U_{I}^{\dagger})

is defined with τa\tau^{a} as an Nc×NcN_{c}\times N_{c} matrix with 2×22\times 2 Pauli matrices embedded in the upper left corner. For the anti-instanton field, we substitute η¯μνa\bar{\eta}^{a}_{\mu\nu} by ημνa\eta^{a}_{\mu\nu} and flip the sign in front of Levi-Cevita tensor, ϵμνρλϵμνρλ\epsilon_{\mu\nu\rho\lambda}\rightarrow-\epsilon_{\mu\nu\rho\lambda}.

The effects of the quark masses on the quark propagator in the instanton or antiinstanton fields, are not known in closed form Liu and Zahed (2021). However, for small masses, the propagator can be expanded around the chiral limit

SI(x,y)=\displaystyle S_{I}(x,y)= x|1i∂̸+I+im|y\displaystyle\langle x|\frac{1}{i\not{\partial}+\not{A}_{I}+im}|y\rangle (35)
=\displaystyle= SZM(I)(x,y)+SNZM(I)(x,y)+𝒪(m)\displaystyle S^{(I)}_{\mathrm{ZM}}(x,y)+S^{(I)}_{\mathrm{NZM}}(x,y)+\mathcal{O}(m)

The zero mode propagator reads

SZM(I)(x,y)=ϕI(x)ϕI(y)imS^{(I)}_{\mathrm{ZM}}(x,y)=\frac{\phi_{I}(x)\phi_{I}^{\dagger}(y)}{im} (36)

where the left-handed zero mode in singular gauge is

ϕI(x)=(ρπ|x|(x2+ρ2)3/2)1γ52UIχ\phi_{I}(x)=\left(\frac{\rho}{\pi|x|(x^{2}+\rho^{2})^{3/2}}\right)\not{x}\frac{1-\gamma^{5}}{2}U_{I}\chi (37)

with χ\chi, a color-spin locked 4-spinor.

The non-zero mode propagator in the chiral-split form reads Brown et al. (1978)

SNZM(I)(x,y)\displaystyle S^{(I)}_{\mathrm{NZM}}(x,y) (38)
=\displaystyle= ixΔI(x,y)1+γ52+ΔI(x,y)iy1γ52\displaystyle i\overrightarrow{\not{D}}_{x}\Delta_{I}(x,y)\frac{1+\gamma^{5}}{2}+\Delta_{I}(x,y)i\overleftarrow{\not{D}}_{y}\frac{1-\gamma^{5}}{2}
=\displaystyle= Snz(I)(x,y)1+γ52+S¯nz(I)(x,y)1γ52\displaystyle S^{(I)}_{nz}(x,y)\frac{1+\gamma^{5}}{2}+\bar{S}^{(I)}_{nz}(x,y)\frac{1-\gamma^{5}}{2}

where ψ(x)Dμ=(μiAμ)ψ(x)\psi(x)\overleftarrow{D}_{\mu}=(-\partial_{\mu}-iA_{\mu})\psi(x). The massless scalar propagator in the single instanton background field is defined as Brown et al. (1978)

ΔI(x,y)=14π2(xy)2(1+ρ2xμyνx2y2UIτμτν+UI)1Π(x)1/2Π(y)1/2\displaystyle\Delta_{I}(x,y)=\frac{1}{4\pi^{2}(x-y)^{2}}\left(1+\rho^{2}\frac{x_{\mu}y_{\nu}}{x^{2}y^{2}}U_{I}\tau_{\mu}^{-}\tau_{\nu}^{+}U_{I}^{\dagger}\right)\frac{1}{\Pi(x)^{1/2}\Pi(y)^{1/2}} (39)

The location of the instanton zIz_{I} is set to be zero for simplicity and can be recovered by translational symmetry xxzIx\rightarrow x-z_{I} and yyzIy\rightarrow y-z_{I}. After a few steps of algebraic calculation, Snz(I)S^{(I)}_{nz} and S¯nz(I)\bar{S}^{(I)}_{nz} can be recast in the form Schäfer and Shuryak (1998); Zubkov et al. (1999); Liu and Zahed (2021)

Snz(I)(x,y)=\displaystyle S^{(I)}_{nz}(x,y)= [i()2π2(xy)4(1+ρ2xμyνx2y2UIτμτν+UI)ρ2γμ4π2xρ(xy)νyλ(x2+ρ2)x2(xy)2y2UIτρτμ+τντλ+UI]\displaystyle\left[\frac{-i(\not{x}-\not{y})}{2\pi^{2}(x-y)^{4}}\left(1+\rho^{2}\frac{x_{\mu}y_{\nu}}{x^{2}y^{2}}U_{I}\tau_{\mu}^{-}\tau_{\nu}^{+}U_{I}^{\dagger}\right)-\frac{\rho^{2}\gamma_{\mu}}{4\pi^{2}}\frac{x_{\rho}(x-y)_{\nu}y_{\lambda}}{(x^{2}+\rho^{2})x^{2}(x-y)^{2}y^{2}}U_{I}\tau_{\rho}^{-}\tau^{+}_{\mu}\tau_{\nu}^{-}\tau_{\lambda}^{+}U_{I}^{\dagger}\right] (40)
×1Π(x)1/2Π(y)1/2\displaystyle\times\frac{1}{\Pi(x)^{1/2}\Pi(y)^{1/2}}

and

S¯nz(I)(x,y)=\displaystyle\bar{S}^{(I)}_{nz}(x,y)= [i()2π2(xy)4(1+ρ2xμyνx2y2UIτμτν+UI)ρ2γμ4π2xρ(xy)νyλ(y2+ρ2)x2(xy)2y2UIτρτν+τμτλ+UI]\displaystyle\left[\frac{-i(\not{x}-\not{y})}{2\pi^{2}(x-y)^{4}}\left(1+\rho^{2}\frac{x_{\mu}y_{\nu}}{x^{2}y^{2}}U_{I}\tau_{\mu}^{-}\tau_{\nu}^{+}U_{I}^{\dagger}\right)-\frac{\rho^{2}\gamma_{\mu}}{4\pi^{2}}\frac{x_{\rho}(x-y)_{\nu}y_{\lambda}}{(y^{2}+\rho^{2})x^{2}(x-y)^{2}y^{2}}U_{I}\tau_{\rho}^{-}\tau^{+}_{\nu}\tau_{\mu}^{-}\tau_{\lambda}^{+}U_{I}^{\dagger}\right] (41)
×1Π(x)1/2Π(y)1/2\displaystyle\times\frac{1}{\Pi(x)^{1/2}\Pi(y)^{1/2}}

Note that the propagator SNZM(A)S^{(A)}_{\mathrm{NZM}} for the anti-instanton can be obtained via the substitutions τμτμ+\tau^{-}_{\mu}\leftrightarrow\tau^{+}_{\mu}, and γ5γ5\gamma^{5}\leftrightarrow-\gamma^{5}.

Appendix B Reduction scheme

The on-shell reduction of the Euclidean and massless quark propagator in the instanton background is subtle. In principle, it can be achieved in two ways: 1/ through LSZ reduction in the zero momentum limit; 2/ through the long term ”time” asymptotic. Here we will quote both procedures, although the long time asymptotics is the one that turns out to be IR free in most cases Liu and Zahed (2021) .

B.1 Zero Euclidean momentum

For massless quarks, the LSZ reduction in the zero momentum (k2Q2k^{2}\ll Q^{2}) limit reads

d4yeikySnz(x,y)ψL(k)eikx(1+ρ2/x2)1/2[1+(1eikx)ρ22x2xμkνkxUτμτν+U]ψL(k)\int d^{4}ye^{-ik\cdot y}S_{nz}(x,y)\not{k}\psi_{L}(k)\simeq\frac{e^{-ik\cdot x}}{(1+\rho^{2}/x^{2})^{1/2}}\left[1+(1-e^{ik\cdot x})\frac{\rho^{2}}{2x^{2}}\frac{x_{\mu}k_{\nu}}{k\cdot x}U\tau^{-}_{\mu}\tau^{+}_{\nu}U^{\dagger}\right]\psi_{L}(k) (42)
d4yeikyψ¯R(k)S¯nz(y,x)eikx(1+ρ2/x2)1/2[1+(1eikx)ρ22x2kμxνkxUτμτν+U]ψ¯R(k)\int d^{4}ye^{ik\cdot y}\bar{\psi}_{R}(k)\not{k}\bar{S}_{nz}(y,x)\simeq\frac{e^{ik\cdot x}}{(1+\rho^{2}/x^{2})^{1/2}}\left[1+(1-e^{-ik\cdot x})\frac{\rho^{2}}{2x^{2}}\frac{k_{\mu}x_{\nu}}{k\cdot x}U\tau^{-}_{\mu}\tau^{+}_{\nu}U^{\dagger}\right]\bar{\psi}_{R}(k) (43)

In the asymptotic limit x2ρ2x^{2}\gg\rho^{2}, the reduction yields an on-shell free quark.

B.2 Large Euclidean time

The alternative reduction scheme that proves to be IR safe in most cases, consists in taking the large Euclidean ”time” asymptotics, to put the quark on mass shell, a common procedure on the lattice. More specifically, we have

limτid3ySnz(x,y)eiky[γ4ψL(k)e|k|τ]\displaystyle\lim_{\tau\rightarrow-\infty}i\int d^{3}\vec{y}S_{nz}(x,y)e^{i\vec{k}\cdot\vec{y}}\left[\gamma_{4}\psi_{L}(k)e^{-|\vec{k}|\tau}\right] (44)
=\displaystyle= eikx(1+ρ2/x2)1/2[1iρ22x2xμkνUτμτν+U01dttei(1t)kx\displaystyle\frac{e^{-ik\cdot x}}{(1+\rho^{2}/x^{2})^{1/2}}\Bigg{[}1-i\frac{\rho^{2}}{2x^{2}}x_{\mu}k_{\nu}U\tau^{-}_{\mu}\tau^{+}_{\nu}U^{\dagger}\int_{0}^{1}dtte^{i(1-t)k\cdot x}
+iρ24x2(x2+ρ2)xμγνγ4|k|xρkλUτμτν+τρτλ+U01dtei(1t)kx+ρ22x2(x2+ρ2)xμγνγ4|k|Uτμτν+U]ψL(k)\displaystyle+i\frac{\rho^{2}}{4x^{2}(x^{2}+\rho^{2})}\frac{x_{\mu}\gamma_{\nu}\gamma_{4}}{|\vec{k}|}x_{\rho}k_{\lambda}U\tau^{-}_{\mu}\tau^{+}_{\nu}\tau^{-}_{\rho}\tau^{+}_{\lambda}U^{\dagger}\int_{0}^{1}dte^{i(1-t)k\cdot x}+\frac{\rho^{2}}{2x^{2}(x^{2}+\rho^{2})}\frac{x_{\mu}\gamma_{\nu}\gamma_{4}}{|\vec{k}|}U\tau^{-}_{\mu}\tau^{+}_{\nu}U^{\dagger}\Bigg{]}\psi_{L}(k)
limτid3y[ψ¯R(k)γ4e|k|τ]S¯nz(y,x)eiky\displaystyle\lim_{\tau\rightarrow\infty}i\int d^{3}\vec{y}\left[\bar{\psi}_{R}(k)\gamma_{4}e^{|\vec{k}|\tau}\right]\bar{S}_{nz}(y,x)e^{-i\vec{k}\cdot\vec{y}} (45)
=\displaystyle= eikx(1+ρ2/x2)1/2ψ¯R(k)[1+iρ22x2kμxνUτμτν+U01dttei(1t)kx\displaystyle\frac{e^{ik\cdot x}}{(1+\rho^{2}/x^{2})^{1/2}}\bar{\psi}_{R}(k)\Bigg{[}1+i\frac{\rho^{2}}{2x^{2}}k_{\mu}x_{\nu}U\tau^{-}_{\mu}\tau^{+}_{\nu}U^{\dagger}\int_{0}^{1}dtte^{-i(1-t)k\cdot x}
iρ24x2(x2+ρ2)γ4γν|k|kλxρxμUτλτρ+τντμ+U01dtei(1t)kx+ρ22x2(x2+ρ2)γ4γμxν|k|Uτμτν+U]\displaystyle-i\frac{\rho^{2}}{4x^{2}(x^{2}+\rho^{2})}\frac{\gamma_{4}\gamma_{\nu}}{|\vec{k}|}k_{\lambda}x_{\rho}x_{\mu}U\tau^{-}_{\lambda}\tau^{+}_{\rho}\tau^{-}_{\nu}\tau^{+}_{\mu}U^{\dagger}\int_{0}^{1}dte^{-i(1-t)k\cdot x}+\frac{\rho^{2}}{2x^{2}(x^{2}+\rho^{2})}\frac{\gamma_{4}\gamma_{\mu}x_{\nu}}{|\vec{k}|}U\tau^{-}_{\mu}\tau^{+}_{\nu}U^{\dagger}\Bigg{]}

where we have used the fact that

12(1σk|k|)ψL(k)=ψL(k)\frac{1}{2}\left(1-\frac{\vec{\sigma}\cdot\vec{k}}{|\vec{k}|}\right)\psi_{L}(k)=\psi_{L}(k)

and

12(1+σk|k|)ψR(k)=ψR(k)\frac{1}{2}\left(1+\frac{\vec{\sigma}\cdot\vec{k}}{|\vec{k}|}\right)\psi_{R}(k)=\psi_{R}(k)

Combined with the reduction scheme in large Euclidean time, (23) reads

V+(k,k)=\displaystyle V_{+}(k^{\prime},k)= 2ρ2Ncmd4xeiqx{eikx(x2+ρ2)218Trc[(1+iρ22x2kσxντστν+01dttei(1t)kx\displaystyle-\frac{2\rho^{2}}{N_{c}m^{*}}\int d^{4}xe^{-iq\cdot x}\Bigg{\{}\frac{e^{ik^{\prime}\cdot x}}{(x^{2}+\rho^{2})^{2}}\frac{1}{8}\mathrm{Tr}_{c}\Bigg{[}\bigg{(}1+i\frac{\rho^{2}}{2x^{2}}k^{\prime}_{\sigma}x_{\nu}\tau^{-}_{\sigma}\tau^{+}_{\nu}\int_{0}^{1}dtte^{-i(1-t)k^{\prime}\cdot x} (46)
iρ24x2(x2+ρ2)γ4γσ|k|kνxρxλτντρ+τστλ+01𝑑tei(1t)kx\displaystyle-i\frac{\rho^{2}}{4x^{2}(x^{2}+\rho^{2})}\frac{\gamma_{4}\gamma_{\sigma}}{|\vec{k}|}k^{\prime}_{\nu}x_{\rho}x_{\lambda}\tau^{-}_{\nu}\tau^{+}_{\rho}\tau^{-}_{\sigma}\tau^{+}_{\lambda}\int_{0}^{1}dte^{-i(1-t)k^{\prime}\cdot x}
+ρ22x2(x2+ρ2)γ4γσxν|k|τστν+)γμγαγβτατβ+]h.c.(kk)}\displaystyle+\frac{\rho^{2}}{2x^{2}(x^{2}+\rho^{2})}\frac{\gamma_{4}\gamma_{\sigma}x_{\nu}}{|\vec{k}|}\tau^{-}_{\sigma}\tau^{+}_{\nu}\bigg{)}\gamma^{\mu}\not{x}\gamma_{\alpha}\gamma_{\beta}\tau^{-}_{\alpha}\tau_{\beta}^{+}\Bigg{]}-~{}\ \begin{matrix}h.c.\\ (k^{\prime}\rightarrow k)\end{matrix}\Bigg{\}}

Appendix C Determinantal mass mm^{*} in the instanton liquid

In the instanton liquid model, the concept of a determinantal mass mm^{*} characterises the width of the zero-mode zone, a key characteristic of the spontaneous breaking of chiral symmetry. It is sampled by the ensemble average of the fermionic determinant

ρNNffDet()ZM=(ρm)NNf\left\langle\rho^{NN_{f}}\prod_{f}\mathrm{Det}(\not{D})_{\mathrm{ZM}}\right\rangle=(\rho m^{*})^{NN_{f}} (47)

a measure of the unquenching of the gauge configurations. In particular, a reduction in the density of pseudoparticles at low resolution

nI+A2=𝑑ρn(ρ)(mρ)Nf\frac{n_{I+A}}{2}=\int d\rho n(\rho)(m^{*}\rho)^{N_{f}} (48)

where n(ρ)n(\rho) denotes the quenched instanton size distribution in (3), and mm^{*} is the determinantal mass. Numerical simulations of ensembles of interacting pseudoparticles give m103MeVm^{*}\sim 103\,\rm MeV Faccioli and Shuryak (2001)

In the thermodynamic limit (N,VN,V\rightarrow\infty with nI+An_{I+A} fixed) alongside the large NcN_{c} limit, the emergent instantonic vertices exponentiate, giving

ZN±𝒟ψ𝒟ψexp(d4xeff)Z_{N_{\pm}}\propto\int\mathcal{D}\psi\mathcal{D}\psi^{\dagger}\exp\left(-\int d^{4}x\mathcal{L}_{\mathrm{eff}}\right) (49)

where the effective Lagrangian in Euclidean space reads Liu et al. (2024)

eff=ψi∂̸ψGI(1+δ)θ+GI(1δ)θ\displaystyle\mathcal{L}_{\mathrm{eff}}=-\psi^{\dagger}i\not{\partial}\psi-G_{I}(1+\delta)\theta_{+}-G_{I}(1-\delta)\theta_{-} (50)

where the vertices are defined as

θ±(x)=𝑑UIf[mf4π2ρ2+iψf(x)UI18τμτν±γμγνUI1γ52ψf(x)]\theta_{\pm}(x)=\int dU_{I}\prod_{f}\left[\frac{m_{f}}{4\pi^{2}\rho^{2}}+i\psi^{\dagger}_{f}(x)U_{I}\frac{1}{8}\tau^{\mp}_{\mu}\tau^{\pm}_{\nu}\gamma_{\mu}\gamma_{\nu}U_{I}^{\dagger}\frac{1\mp\gamma^{5}}{2}\psi_{f}(x)\right] (51)

The emergent parameters GIG_{I} and δ\delta are fixed by the saddle point approximation. The effective coupling GIG_{I}

GI=N2V(4π2ρ2m)Nf\displaystyle G_{I}=\frac{N}{2V}\left(\frac{4\pi^{2}\rho^{2}}{m^{*}}\right)^{N_{f}} (52)

is tied to the mean instanton size ρ\rho, density N/VN/V, and determinantal mass mm^{*} Schäfer and Shuryak (1996); Faccioli and Shuryak (2001); Shuryak and Zahed (2023); Liu et al. (2024) The screened topological charge δ\delta is

δ=NfmmΔN\delta=N_{f}\frac{m^{*}}{m}\frac{\Delta}{N} (53)

In the saddle point approximation, the momentum dependent constituent mass is

M(k)=N2Vk2φ(k)2mM(k)=\frac{N}{2V}\frac{k^{2}\varphi^{\prime}(k)^{2}}{m^{*}} (54)

along with the gap equation which naturally determines the determinantal mass Liu et al. (2024)

m=m+8π2ρ2d4k(2π)4M(k)(k)k2+M2(k)m^{*}=m+8\pi^{2}\rho^{2}\int\frac{d^{4}k}{(2\pi)^{4}}\frac{M(k)\mathcal{F}(k)}{k^{2}+M^{2}(k)} (55)

where the instanton-quark form factor is defined as

(k)=zddz[I0(z)K0(z)I1(z)K1(z)]|z=k2\sqrt{\mathcal{F}(k)}=z\frac{d}{dz}[I_{0}(z)K_{0}(z)-I_{1}(z)K_{1}(z)]\bigg{|}_{z=\frac{k}{2}} (56)

Using the instanton parameters nI+A=1n_{I+A}=1 fm-4, ρ=0.313\rho=0.313 fm and the current mass m=8m=8 MeV, the determinantal mass is

m=102.1MeV\displaystyle m^{*}=102.1~{}\mathrm{MeV} (57)

with the constituent mass of 395395 MeV Liu et al. (2023a, b); Liu (2025), and a quark condensate (275MeV)3-(275\mathrm{MeV})^{3}. The estimated determinantal mass is much closer to the numerical value of m103m^{*}\simeq 103 MeV Faccioli and Shuryak (2001).

Appendix D Averaging in grand canonical instanton liquid ensemble

Here we briefly outline the averaging over the fluctuations in the number of pseudoparticles in the ILM. In a grand canonical description where N±N_{\pm} are allowed to fluctuate, the partition function of the QCD vacuum with finite vacuum angle is written as

𝒵(μ,θ)=\displaystyle\mathcal{Z}(\mu,\theta)= N+,NZN±e(μ+iθ)N++(μiθ)N\displaystyle\sum_{N_{+},N_{-}}Z_{N_{\pm}}e^{(\mu+i\theta)N_{+}+(\mu-i\theta)N_{-}} (58)
=\displaystyle= exp(b4Nθe4μb)\displaystyle\exp\left(\frac{b}{4}\langle N\rangle_{\theta}e^{\frac{4\mu}{b}}\right)

This partition function yield the distribution with the measures (6) and (8) at μ=0\mu=0 and θ=0\theta=0 or equivalently written as Diakonov et al. (1996); Schäfer and Shuryak (1998); Zahed (2021)

𝒫(N+,N)=(N)(Δ)\mathds{\mathcal{P}}(N_{+},N_{-})=\mathbb{P}(N)\mathbb{Q}(\Delta) (59)

with mean N¯=N\bar{N}=\langle N\rangle and Qt=ΔQ_{t}=\langle\Delta\rangle. As a result, vacuum expectation values of most quarks and gluon operators are averaged through

𝒪=N+,N𝒫(N+,N)𝒪N±𝒪¯N±\langle\mathcal{O}\rangle=\sum_{N_{+},N_{-}}\mathcal{P}(N_{+},N_{-})\langle\mathcal{O}\rangle_{N_{\pm}}\equiv\overline{\langle\mathcal{O}\rangle}_{N_{\pm}} (60)

The averaging is carried out over the configurations with fixed N±N_{\pm} (canonical ensemble average), followed by an ensemble averaging over the distribution (8).

Similarly, the evaluation of the hadronic matrix elements can be formulated as a large-TT reduction of a 3-point function

h|𝒪|hh|h=limTJh(T/2)𝒪Jh(T/2)conJh(T/2)Jh(T/2)\frac{\langle h|\mathcal{O}|h\rangle}{\langle h|h\rangle}=\lim_{T\rightarrow\infty}\frac{\langle J^{\dagger}_{h}(T/2)\mathcal{O}J_{h}(-T/2)\rangle_{\mathrm{con}}}{\langle J^{\dagger}_{h}(T/2)J_{h}(-T/2)\rangle} (61)

where Jh(t)J_{h}(t) is a pertinent source for the hadronic state hh.

Now with this grand canonical framework, the insertion of the gluonic scalar operator and topological charge operator in the the connected correlation function can be carried out by

132π2𝒪d4xFμνFμνcon=\displaystyle\frac{1}{32\pi^{2}}\left\langle\mathcal{O}\int d^{4}xF_{\mu\nu}F_{\mu\nu}\right\rangle_{\mathrm{con}}= (N2N2)N𝒪N±|N=N¯Δ=0\displaystyle\left(\langle N^{2}\rangle-\langle N\rangle^{2}\right)\frac{\partial}{\partial N}\left\langle\mathcal{O}\right\rangle_{N_{\pm}}\bigg{|}_{\begin{subarray}{c}N=\bar{N}\\ \Delta=0\end{subarray}} (62)
132π2𝒪d4xFμνF~μνcon=\displaystyle\frac{1}{32\pi^{2}}\left\langle\mathcal{O}\int d^{4}xF_{\mu\nu}\tilde{F}_{\mu\nu}\right\rangle_{\mathrm{con}}= Δ2Δ𝒪|N=N¯Δ=0\displaystyle\langle\Delta^{2}\rangle\frac{\partial}{\partial\Delta}\left\langle\mathcal{O}\right\rangle\bigg{|}_{\begin{subarray}{c}N=\bar{N}\\ \Delta=0\end{subarray}} (63)

The quark contribution are usually penalized by 1/Nc1/N_{c}-counting, as they are rooted in the quark-instanton interaction. Therefore, in this case, the leading contribution will come from the fluctuations.

This approach can used to calculate the neutron EDM.

dN=N(θ)|d3xxJ0(x)|N(θ)\displaystyle\vec{d}_{N}=\langle N(\theta)|\int d^{3}\vec{x}\,\vec{x}J_{0}(\vec{x})|N(\theta)\rangle (64)
iθN(0)|d3xxJ0(x)(132π2d4zFμνF~μν)|N(0)\displaystyle\simeq-i\theta\langle N(0)|\int d^{3}\vec{x}\,\vec{x}J_{0}(\vec{x})\left(\frac{1}{32\pi^{2}}\int d^{4}z\,F_{\mu\nu}\tilde{F}_{\mu\nu}\right)|N(0)\rangle

In the 1/Nc1/N_{c} book-keeping, the dominant contributions to neutron EDM are given by

dN=iθΔ2ΔN|d3xxJ0(x)|N|N=N¯Δ=0\displaystyle\vec{d}_{N}=-i\theta\langle\Delta^{2}\rangle\frac{\partial}{\partial\Delta}\langle N|\int d^{3}\vec{x}\,\vec{x}J_{0}(\vec{x})|N\rangle\bigg{|}_{\begin{subarray}{c}N=\bar{N}\\ \Delta=0\end{subarray}} (65)

References