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

Computing Nucleon Electric Dipole Moment from lattice QCD

Taku Izubuchia,b, a,c, Sergey Syritsyna,d,
a RIKEN/BNL Research Center, Brookhaven National Laboratory, Upton, NY 11973, USA
b Physics Department, Brookhaven National Laboratory, Upton, New York 11973, USA
c Department of Physics, Nara Women’s University, Nara 630-8506, Japan
d Department of Physics and Astronomy, Stony Brook University, Stony Brook, NY 11794, USA
E-mail
Abstract:

Electric dipole moments (EDMs) of nucleons and nuclei are actively considered as direct evidence of the CP violation. Calculations of nucleon EDMs on lattice are required to connect the quark- and hadron- level effective CP violating interactions within QCD or other CP violating sources in new physics beyond the standard model. Among them, the theta-induced nucleon EDM, that is the only such renormalizable interaction, has widely been investigated on a lattice. In the report, we review recent developments of the lattice calculations of nucleon EDM induced QCD theta term.

1 Introduction

Observation of a non-zero nucleon electric dipole moment (nEDM) would be direct evidence for violation of CP(T)-symmetry. The Standard Model (SM) has a CP-violating phase in the CKM matrix. However, there is at least one substantial motivation to search for CP violation beyond the SM: the magnitude of the CP violation effect in CKM is insufficient to explain the observed baryon asymmetry of the universe as required by one of the famous Sakharov’s conditions for the universe’s matter origin. Currently the most sensitive probes for the CP-violating phenomena are EDM searches in hadronic, atomic, and molecular systems. The best limits on EDMs come from experiments on neutrons (ILL) [1] and 199Hg [2] which constrain the nucleon EDM (nEDM) |dn|2.6×1026|d_{n}|\leq 2.6\times 10^{-26} [ee cm]. This bound is 105\sim 10^{5} larger than the prediction 1031\sim 10^{-31} [ee cm] from the CP phase of the CKM matrix in the SM. Observation of non-zero nEDMs would be a discovery of fundamental importance, and however, even a null result would make serious impact on cosmology and high-energy theory. Violations of CP symmetry at the quark level are represented by a number of effective quark and gluon operators. Among them, the only such renormalizable interaction is the QCD ”θ\theta-term”,

iSθ=iθQ=iθxq(x),q(x)=116π2Tr[GμνG~μν(x)]\displaystyle iS_{\theta}=i\theta Q=i\theta\sum_{x}q(x)\,,\quad\quad q(x)=\frac{1}{16\pi^{2}}{\rm Tr}\left[G_{\mu\nu}\tilde{G}^{\mu\nu}(x)\right] (1)

where QQ is the topological charge and q(x)q(x) is the topological charge density operator. Assuming that the θ\theta-term is the only source of the CP violation, we have a strong constraint on the parameter |θ|1010|\theta|\lesssim 10^{-10}. The problem of why θ\theta is so small is known as the strong CP problem 111A dynamical solution to the strong CP problem is proposed in a parallel talk by G. Schierholz [3].. The precision of EDM measurements of nucleons and nuclei will increase in the future experiments using neutron sources, which plan to improve neutron EDMs bounds by 1-2 orders of magnitude. However, quantitative connection between magnitudes of EDMs and such CP violation operators at the quark-gluon level is very limited and model-dependent (see Refs. [4], [5] for a recent review of the EDM phenomenology). Therefore connecting the quark- and hadron-level effective interactions that include CP violating sources is an urgent task for lattice QCD.

In this proceedings, we review recent progress on the lattice calculations of the nEDM induced by the θ\theta term. As for other CP-violating matrix elements that arise from the operators beyond the SM, these are discussed in the plenary lecture [6] 222See also [7] for a recent review..

2 CP odd Form factor and parity mixing

In this section, we briefly review the methods for the lattice computations of the nEDM. The θ\theta-induced nEDMs have been calculated on a lattice from the CP-odd electric dipole form factor (EDFF) F3(Q2)F_{3}(Q^{2}) with the Q20Q^{2}\to 0 extrapolation of the nucleon matrix elements of the quark vector current [8, 9, 10, 11, 12, 13, 14, 15, 16, 17], and from nucleon energy shifts in a uniform background electric field [18, 19, 14]. The EDFF is defined as

p,σ|Jμ|p,σCP=u¯p,σ[F1(Q2)γμ+(F2(Q2)+iF3(Q2))iσμνqν2MN]up,σ,\langle p^{\prime},\sigma^{\prime}|J^{\mu}|p,\sigma\rangle_{{\cancel{\rm CP}}}=\bar{u}_{p^{\prime},\sigma^{\prime}}\big{[}F_{1}(Q^{2})\gamma^{\mu}+\big{(}F_{2}(Q^{2})+iF_{3}(Q^{2})\big{)}\frac{i\sigma^{\mu\nu}q_{\nu}}{2M_{N}}\big{]}u_{p,\sigma}\,, (2)

where Q2=q2Q^{2}=-q^{2} and q=ppq=p^{\prime}-p, and F1F_{1} and F2F_{2} are the Dirac and Pauli form factors. The QCD θ\theta term is introduced either as a modification of the lattice CP-even QCD action with imaginary θ\theta term [10, 11] or as a Taylor expansion of nucleon correlation functions, SQCDSQCD+iSθS_{\text{QCD}}\to S_{\text{QCD}}+iS_{\theta}. In the latter case the nucleon-current correlation functions in CP{\cancel{\rm CP}} QCD vacuum are modified as

N[q¯γμq]N¯CP=1Z𝒟U𝒟ψ¯𝒟ψN[q¯γμq]N¯eSiSθ=CNJN¯iθCNJN¯Q+O(θ2),\displaystyle\begin{aligned} \langle N\,[\bar{q}\gamma^{\mu}q]\,\bar{N}\rangle_{{\cancel{\rm CP}}}&=\frac{1}{Z}\int\,{\mathcal{D}}U\,\mathcal{D}\bar{\psi}\mathcal{D}\psi N\,[\bar{q}\gamma^{\mu}q]\,\bar{N}e^{-S-iS_{\theta}}\,\\ &=C_{NJ\bar{N}}-i\theta\,C_{NJ\bar{N}}^{Q}+O(\theta^{2})\,,\end{aligned} (3)

where CNJN¯=N[q¯γμq]N¯C_{NJ\bar{N}}=\langle N\,[\bar{q}\gamma^{\mu}q]\,\bar{N}\rangle and CNJN¯Q=N[q¯γμq]N¯x[q(x)]C_{NJ\bar{N}}^{Q}=\langle N\,[\bar{q}\gamma^{\mu}q]\,\bar{N}\,\sum_{x}[q(x)]\rangle are the nucleon-current correlation functions evaluated in the CP{CP}-even QCD vacuum. To obtain the EDFF F3(Q2)F_{3}(Q^{2}) in Eq. (2) we have to calculate the nucleon two and three point functions

C2ptQ(p,t)\displaystyle C_{\rm 2pt}^{Q}(\vec{p},t) =yeipyN(y,t)N¯(0,0)x[q(x)],\displaystyle=\sum_{\vec{y}}e^{-i\vec{p}\cdot\vec{y}}\langle N(\vec{y},t)\bar{N}(\vec{0},0)\sum_{x}[q(x)]\rangle, (4)
C3ptQ(t,p;top,q)\displaystyle C_{\rm 3pt}^{Q}(t,\vec{p};t_{\rm op},\vec{q}) =y,zeipy+qzN(y,t)Jμ(z,top)N¯(0,0)x[q(x)].\displaystyle=\sum_{\vec{y},\vec{z}}e^{-i\vec{p}\cdot\vec{y}+\vec{q}\cdot\vec{z}}\langle N(\vec{y},t)J^{\mu}(\vec{z},t_{\rm op})\bar{N}(\vec{0},0)\sum_{x}[q(x)]\rangle. (5)

In general the nucleon ground states as well as their overlaps with the positive-parity nucleon ground state are modified in CP{\cancel{\rm CP}} QCD vacuum as

0|N|p,σCP=ZNeiαγ5up,σ=ZNu~p,σ,\displaystyle\langle 0|N|p,\sigma\rangle_{\cancel{\rm CP}}=Z_{N}e^{i\alpha\gamma_{5}}u_{p,\sigma}=Z_{N}\tilde{u}_{p,\sigma}, (6)

where u~p,σ\tilde{u}_{p,\sigma} is a spinor wave function for the nucleon state |p,σCP|p,\sigma\rangle_{\cancel{\rm CP}} and ZNZ_{N} is a normalization constant. The spinor u~p,σ\tilde{u}_{p,\sigma} satisfies the following free Dirac equation with CP-violating γ5\gamma_{5} mass

(/pmNe2iαγ5)u~p,σ=(/pmNe2iαγ5)eiαγ5up,σ=0,\displaystyle({\ooalign{\hfil/\hfil\crcr$p$}}-m_{N}e^{-2i\alpha\gamma_{5}})\tilde{u}_{p,\sigma}=({\ooalign{\hfil/\hfil\crcr$p$}}-m_{N}e^{-2i\alpha\gamma_{5}})e^{i\alpha\gamma_{5}}u_{p,\sigma}=0, (11)

where up,σu_{p,\sigma} is a wave function spinor in CP-even QCD vacuum. Thus when the CP violating effect exists in the QCD vacuum, one should use the modified nucleon spinor u~p,σ\tilde{u}_{p,\sigma} in lattice calculations which affects the kinematic coefficients. For example, ignoring excite states, the nucleon two-point correlation function with CP violating operator can be represented as

C2ptQ(p,t)\displaystyle C_{\rm 2pt}^{Q}(\vec{p},t) =|ZN|2eEpt2Epσu~p,σu~¯p,σ=|ZN|2eEpt2Ep(mNe2iαγ5i/p),\displaystyle=|Z_{N}|^{2}\frac{e^{-E_{p}t}}{2E_{p}}\sum_{\sigma}\tilde{u}_{p,\sigma}\bar{\tilde{u}}_{p,\sigma}=|Z_{N}|^{2}\frac{e^{-E_{p}t}}{2E_{p}}\left(m_{N}e^{2i\alpha\gamma_{5}}-i{\ooalign{\hfil/\hfil\crcr$p$}}\right), (14)

where we use the completeness condition for the free Dirac spinor,

σu~p,σu~¯p,σ=eiαγ5(σup,σu¯p,σ)=mNe2iαγ5i/p.\displaystyle\sum_{\sigma}\tilde{u}_{p,\sigma}\bar{\tilde{u}}_{p,\sigma}=e^{i\alpha\gamma_{5}}\left(\sum_{\sigma}u_{p,\sigma}\bar{u}_{p,\sigma}\right)=m_{N}e^{2i\alpha\gamma_{5}}-i{\ooalign{\hfil/\hfil\crcr$p$}}. (17)

The modification also affects the nucleon matrix elements p,σ|Jμ|p,σCP\langle p^{\prime},\sigma^{\prime}|J^{\mu}|p,\sigma\rangle_{\cancel{\rm CP}} as

u~¯p,σ[F1~γμ+(F2~+iF3~γ5)iσμνqν2mN]u~p,σ\displaystyle\bar{\tilde{u}}_{p^{\prime},\sigma^{\prime}}\left[\tilde{F_{1}}\gamma^{\mu}+(\tilde{F_{2}}+i\tilde{F_{3}}\gamma_{5})\frac{i\sigma^{\mu\nu}q_{\nu}}{2m_{N}}\right]\tilde{u}_{p,\sigma} =u¯p,σ[F1~γμ+e2iαγ5(F2~+iF3~γ5)iσμνqν2mN]up,σ\displaystyle=\bar{u}_{p^{\prime},\sigma^{\prime}}\left[\tilde{F_{1}}\gamma^{\mu}+e^{2i\alpha\gamma_{5}}(\tilde{F_{2}}+i\tilde{F_{3}}\gamma_{5})\frac{i\sigma^{\mu\nu}q_{\nu}}{2m_{N}}\right]u_{p,\sigma}
=u¯p,σ[F1γμ+(F2+iF3γ5)iσμνqν2mN]up,σ.\displaystyle=\bar{u}_{p^{\prime},\sigma^{\prime}}\left[F_{1}\gamma^{\mu}+(F_{2}+iF_{3}\gamma_{5})\frac{i\sigma^{\mu\nu}q_{\nu}}{2m_{N}}\right]u_{p,\sigma}.

In the original lattice calculation [8], while the CP violation effect on the kinematic coefficient in Eq. (17) have been correctly taken into account, the inadequate definition of these form factors of F~1,2,3\tilde{F}_{1,2,3} has been used prior to [14]. As a result, all previous lattice results has a spurious contributions to the EDFF computed with F~3\tilde{F}_{3} from the Pauli form factor F2F_{2} as

F3~=F32αF2.\displaystyle\tilde{F_{3}}=F_{3}-2\alpha F_{2}. (18)

Thus if the ”Old” formula (F3~\tilde{F_{3}}) is used for extracting the nEDM d~n=F~3(0)/(2mN)\tilde{d}_{n}=\tilde{F}_{3}(0)/(2m_{N}), there is a correction from the spurious mixing, which becomes significant when α\alpha becomes large.

The inconsistency between ”Old” and ”New” formula can be directly and numerically confirmed by comparing with computing nEDM from the energy shift method. The uniform electric field that preserves translation invariance and the (anti-) periodic boundary conditions on a lattice was first introduced in [20] to study the nEDM, and also applied to CP-even quantities such as electric polarizability and magnetic moments of the nucleon [21, 22]. The uniform background electric field is analytically continued to an imaginary value, so that the nucleon energy shift due to nEDM becomes imaginary. Expanding the two-point function up to the first order in θ\theta we can directly extract the nEDM contribution that is linear in tt. For simplicity we only consider the neutral particles, since the correlation function of charged particles is more complicated. To introduce the background electric field we consider the following Euclidean U(1)U(1) vector potential

Ax,j=nijΦijxi,Ax,i|xi=Li1=nijΦijLixj,\displaystyle A_{x,j}=n_{ij}\Phi_{ij}x_{i},\quad\quad A_{x,i}|_{x_{i}=L_{i}-1}=-n_{ij}\Phi_{ij}L_{i}x_{j}, (19)

where Φμν=6πLμLν\Phi_{\mu\nu}=\frac{6\pi}{L_{\mu}L_{\nu}} is the quantum field flux on a plaquette (μν\mu\nu) and nμνn_{\mu\nu} is the corresponding number of quanta. The (anti-) periodicity on a lattice in both space and time requires the Dirac quantization conditions

QqΦμνLμLν=2πnμν,(Qu=23,Qd=13)\displaystyle Q_{q}\Phi_{\mu\nu}L_{\mu}L_{\nu}=2\pi n_{\mu\nu},\quad\quad\left(Q_{u}=\frac{2}{3},\ \ Q_{d}=-\frac{1}{3}\right) (20)

The electric field vector is then given as E=(n14Φ14,n24Φ24,n34Φ34)\vec{E}=(n_{14}\Phi_{14},n_{24}\Phi_{24},n_{34}\Phi_{34}). The corresponding effective Dirac equation for nucleon field N~\tilde{N} is given as

(/p(κ~+iζ~γ5)Fμν2σμν2mNmNe2iαγ5)N~=0,\displaystyle\left({\ooalign{\hfil/\hfil\crcr$p$}}-(\tilde{\kappa}+i\tilde{\zeta}\gamma_{5})\frac{F_{\mu\nu}}{2}\frac{\sigma^{\mu\nu}}{2m_{N}}-m_{N}e^{-2i\alpha\gamma_{5}}\right)\tilde{N}=0, (23)

where κ~=F~2(0)\tilde{\kappa}=\tilde{F}_{2}(0) and ζ~=F~3(0)\tilde{\zeta}=\tilde{F}_{3}(0) are the effective anomalous magnetic and electric moments in the basis of N~\tilde{N} with γ5\gamma_{5} mass. It is obvious that the CP{\cancel{\rm CP}} phase eiαγ5e^{i\alpha\gamma_{5}} can be completely rotated away by a field redefinition N=eiαγ5N~N=e^{i\alpha\gamma_{5}}\tilde{N}, where the two couplings also transform e2iα(κ~+iζ~)=(κ+iζ)e^{2i\alpha}(\tilde{\kappa}+i\tilde{\zeta})=(\kappa+i\zeta). Considering an electric field in zz-direction E=(0,0,Ez)\vec{E}=(0,0,E_{z}) in the rest frame pμ=(0,Es)p^{\mu}=(\vec{0},E_{s}), we obtain an on-shell solution spinor uEz,su_{E_{z},s} for the spin polarized along zz-direction with s=±s=\pm, which has a spin-dependent energy eigenvalue Es=mNζ2mN(sEz)+O(Ez2)E_{s}=m_{N}-\frac{\zeta}{2m_{N}}(sE_{z})+O(E_{z}^{2}). From the result we see that the resulting energy shift is consistent with the EDFF obtained in the basis of NN without γ5\gamma_{5} mass, and κ=F2(0)\kappa=F_{2}(0) and ζ=F3(0)\zeta=F_{3}(0) are the nucleon magnetic and electric dipole moment coefficients. By taking the analytic continuation of the electric field EzE_{z} to the imaginary value we obtain the energy shift of a nucleon on lattice as E~±=mN±δE\tilde{E}_{\pm}=m_{N}\pm\delta E, with δE=ζ/(2mN)iEz\delta E=-\zeta/(2m_{N})iE_{z}. To extract the nucleon energy shift we calculate the following nucleon two point functions in the presence of the background electric fields

C2pt,EQ(p,t)=yeipyN(y,t)|N¯(0,0)x[q(x)]E.\displaystyle C_{{\rm 2pt},\vec{E}}^{Q}(\vec{p},t)=\sum_{\vec{y}}e^{-i\vec{p}\cdot\vec{y}}\langle N(\vec{y},t)|\bar{N}(\vec{0},0)\sum_{x}[q(x)]\rangle_{\vec{E}}. (24)

Using the solution spinor uEz,su_{E_{z},s} and ignoring excited states we can expand C2pt,EQ(p,t)C_{{\rm 2pt},\vec{E}}^{Q}(\vec{p},t) as

C2pt,EQ(0,t)\displaystyle C_{{\rm 2pt},\vec{E}}^{Q}(\vec{0},t) =|ZN|2s=±u~Ez,su~¯Ez,seE~st2E~s\displaystyle=|Z_{N}|^{2}\sum_{s=\pm}\tilde{u}_{E_{z},s}\bar{\tilde{u}}_{E_{z},s}\frac{e^{-\tilde{E}_{s}t}}{2\tilde{E}_{s}}
=|ZN|2[1+γ42(1ΣzδEt)+iα5γ5+ΣzκEz2m2γ5]emNt+O(δE2,Ez2),\displaystyle=|Z_{N}|^{2}\left[\frac{1+\gamma_{4}}{2}(1-\Sigma^{z}\delta Et)+i\alpha_{5}\gamma_{5}+\Sigma^{z}\frac{\kappa E_{z}}{2m^{2}}\gamma_{5}\right]e^{-m_{N}t}+O(\delta E^{2},E_{z}^{2}), (25)

where Σz=iγxγy\Sigma_{z}=-i\gamma_{x}\gamma_{y}. Using the standard ratio method we define an “effective” energy shift ζneff(t)=2mNdn=2mNEz(R(t+1)R(t))\zeta_{n}^{\rm eff}(t)=2m_{N}d_{n}=\frac{-2m_{N}}{E_{z}}(R(t+1)-R(t)), and R(t)R(t) is defined as a ratio of two-point functions with two different spin projections

R(t)=Tr[(TSz+TSz)C2pt,ECP(0,t)]2Tr[TpC2pt,ECP(0,t)],\displaystyle R(t)=\frac{{\rm Tr}[(T_{S_{z+}}-T_{S_{z-}})C_{{\rm 2pt},\vec{E}}^{\cancel{\rm CP}}(\vec{0},t)]}{2{\rm Tr}[T_{p}C_{{\rm 2pt},\vec{E}}^{\cancel{\rm CP}}(\vec{0},t)]}, (26)

where we use spin polarization projection operators TSz±=1+γ42(1±Σz)T_{S_{z\pm}}=\frac{1+\gamma_{4}}{2}(1\pm\Sigma_{z}), and Tp=(1+γ4)T_{p}=(1+\gamma_{4}). Thus the nEDM in the energy shift method is independent from the parity mixing ambiguity, from which we numerically verify the consistency with the “New” formula (F3F_{3}). For more detail on the analyses, see Ref. [14].

We show some nEDM results in comparison between the form factors and the energy shift methods. Fig. 2 shows how the spurious mixing affects the result for the θ\theta-EDM. As shown in the figure, the magnitudes of F3(Q2)F_{3}(Q^{2}) with ”New” are smaller than the ”Old” values. We also show the result for nEDM ζneff\zeta_{n}^{\rm eff} computed from the effective energy shift in Fig. 2, where we computed with two values of flux quanta n=n34=±1n=n_{34}=\pm 1 and ±2\pm 2. A plateau for the effective energy ζneff\zeta_{n}^{\rm eff} at t=47t=4\sim 7 can be obtained in both electric fields of |n|=1|n|=1 and 22. Comparing the form factor method we obtain a consistent result only if we use the “New” formula. This result is also consistent with naively corrected data using the previously reported values prior to [14]. Since we obtain a small but non-zero signal |dnθ|0.1e2mN|d_{n}^{\theta}|\lesssim 0.1\frac{e}{2m_{N}} at mπ=330m_{\pi}=330 MeV, we estimate an extrapolated value to the physical point based on a naive scaling of the ChPT expectation dnmqmπ2d_{n}\sim m_{q}\sim m_{\pi}^{2}, which suggests that the signal becomes weaker as the quark mass is approaching to the physical point and we would obtain a smaller value of |dnθ(mπphys)|0.02e2mN|d_{n}^{\theta}(m_{\pi}^{\rm phys})|\lesssim 0.02\frac{e}{2m_{N}}. From our findings in order to promote a direct calculation of nEDM at the physical point, various noise reduction techniques that work in particular for gluon operator are required in addition to a significant increase in statistics.

Refer to caption
Figure 1: Plateaus for the nucleon EDFF F3(Q2)F_{3}(Q^{2}) from QCD θ\theta term for top=8t_{op}=8. The results with ”Old” include spurious mixing with the F2F_{2}. Results are shown for a lattice ensemble of domain wall fermion configurations of 243×6424^{3}\times 64 for mπ=330m_{\pi}=330 MeV.
Refer to caption
Figure 2: Effective nEDM ζneff\zeta_{n}^{\rm eff} from QCD θ\theta term. Results are shown for a lattice ensemble of domain wall fermion configurations of 243×6424^{3}\times 64 for mπ=330m_{\pi}=330 MeV.

3 Noise reduction technique for θ\theta-EDM

As explained in the previous section, θ\theta-induced nEDM would be extremely challenging in the physical point, because the topological charge fluctuation dominates the large statistical noise growing with lattice volume V4V_{4} as δQ2=Q2V4\delta Q^{2}=\langle Q^{2}\rangle\propto V_{4}, while the signal becomes small. Since the global topological charge is zero on average because our QCD action is CP-even, F3(Q2)F_{3}(Q^{2}) is the signal of the correlation between the gluon operator and the fermionic (nucleon) functions as shown in Eq. (5). Thus it has been originally suggested that truncation of the topological charge sum at a large distance from the nucleon position can reduce fluctuations of the nEDM [13] (also known as a cluster decomposition of disconnected diagrams [23]), since contributions of QQ at large distance may be neglected in computing the nEDM, while its correlation has a large noise which is not suppressed with space-time distance due to the global nature of the topological charge. To extend this method, in Ref. [24] we consider a generalized reduced topological charge density which separately restrict time and space to a cylindrical volume VQV_{Q},

Q~(ΔtQ,rQ)=116π2xVQTr[G^μνG^~μν]x,(x,t)VQ:{|xx0|rQ,t0ΔtQ<t<t0+tsep+ΔtQ,\displaystyle\tilde{Q}(\Delta t_{Q},r_{Q})=\frac{1}{16\pi^{2}}\sum_{x\in V_{Q}}{\rm Tr}\big{[}\hat{G}_{\mu\nu}\tilde{\hat{G}}_{\mu\nu}\big{]}_{x}\,,\quad(\vec{x},t)\in V_{Q}:\left\{\begin{array}[]{l}|\vec{x}-\vec{x}_{0}|\leq r_{Q}\,,\\ t_{0}-\Delta t_{Q}<t<t_{0}+t_{\rm sep}+\Delta t_{Q}\,,\end{array}\right. (29)

where t0t_{0} and t0+tsept_{0}+t_{\rm sep} are the positions of the nucleon source and sink. This setup is illustrated in Fig. 3, where the three-point functions are inside entirely the region VQV_{Q} and the truncation in tt-direction for Q~\tilde{Q} is symmetric with respect to the nucleon sources and sinks for both two- and three-point functions. We also set x0\vec{x}_{0} to the nucleon source position to further reduce the noise at large distances. We should note that a spatial restriction may introduce another bias for nEDM. In order to avoid such ambiguity, the convergence with rQr_{Q} and ΔtQ\Delta t_{Q} in Eq. (29) must be verified at each nucleon momenta, especially in computing the Q2Q^{2} dependence of the EDFFs.

Refer to caption
Figure 3: Truncated sampling of the topological charge density (29) for reducing the noise in the CP-odd nucleon correlation functions (4) and (5). The correlation with the points outside VQV_{Q} is expected to be suppressed but gives a large noise.

We use the lattice QCD ensemble of dynamical domain wall fermion with heavy pion mass mπ=330m_{\pi}=330 MeV. We calculate 64 low-precision and 1 high-precision samples using the AMA sampling method [25]. The topological charge density is calculated using the “5-loop-improved” field strength G^μν\hat{G}_{\mu\nu} [26] with the gradient flow (tgf=8a2t_{gf}=8a^{2}). The rQr_{Q} and ΔtQ\Delta t_{Q} dependence of the mixing angle α\alpha and the EDFF for the neutron (connected diagram only) are shown in Figs. 4 and 5, where we observe error reduction for smaller values of rQr_{Q} and ΔQ\Delta_{Q} and convergence for rQ16r_{Q}\geq 16, ΔtQ8\Delta t_{Q}\geq 8 in both α\alpha and F3(Q2)F_{3}(Q^{2}). We have also performed a calculation using ensembles at the physical point on 483×9648^{3}\times 96 lattice with 33,000 statistics. Unfortunately we have found no signal for neutron EDFFs (See Fig. 6), and the results are consistent with zero with the statistical uncertainty. Our result of signal-to-noise ratio 0.2\sim 0.2 at mπ=330m_{\pi}=330 MeV indicates that the expected signal-to-noise ratio at the physical point has to be improved by a factor of \sim10 which requires ×O(100)\times O(100) more statistics to confirm the existence of the strong CP problem.

Refer to caption
Figure 4: Dependence of the spatial and temporal cuts (rQr_{Q}, ΔtQ\Delta t_{Q}) in the reduced topological charge (29) on the nucleon parity mixing angle α5\alpha_{5}.
Refer to caption
Figure 5: Neutron EDFF induced by θ\theta-term and its dependence of the spatial and temporal cuts (rQr_{Q}, ΔtQ\Delta t_{Q}) at mπ=330m_{\pi}=330 MeV.
Refer to caption
Figure 6: Preliminary results for the proton and neutron EDFF induced by θ\theta-term and its dependence of the spatial and temporal cuts (rQr_{Q}, ΔtQ\Delta t_{Q}) at mπ=139m_{\pi}=139 MeV [24].

A systematic analysis of the truncation method based on spectrum decomposition is proposed in Refs. [15, 16]. This study uses a truncation of the topological charge density operator summed over spatial directions Q¯\bar{Q} defined as

Q¯(tQ)=d3xq(x,tQ),q(x,tQ)=116π2GG~(x,tQ)\displaystyle\bar{Q}(t_{Q})=\int d^{3}xq(x,t_{Q}),\quad\quad q(x,t_{Q})=\frac{1}{16\pi^{2}}G\tilde{G}(x,t_{Q}) (30)

where the topological charge density q(x,tQ)q(x,t_{Q}) is again calculated using the gradient flow method. To calculate the nucleon mixing angle, the following modified two-point function

ΔC2pt(t,tQ)=T{N(T)Q¯(tQ)N¯(0)},\displaystyle\Delta C_{\rm 2pt}(t,t_{Q})=\langle{\rm T}\{N(T)\bar{Q}(t_{Q})\bar{N}(0)\}\rangle, (31)

is considered. The spectrum decomposition for ΔC2pt(t,tQ)\Delta C_{\rm 2pt}(t,t_{Q}) reveals its tQt_{Q} dependence and provide a systematic way to estimate the truncation error. For example, in the case of 0<tQ<t0<t_{Q}<t, its spectral decomposition has the form

ΔC2pt(t,tQ)=N(t)Q¯(tQ)N¯(0)\displaystyle\Delta C_{\rm 2pt}(t,t_{Q})=\langle N(t)\bar{Q}(t_{Q})\bar{N}(0)\rangle n,meEn(ttQ)EmtQ0|N|nn|Q¯|mm|N¯|0\displaystyle\sim\sum_{n,m}e^{-E_{n}(t-t_{Q})-E_{m}t_{Q}}\langle 0|N|n\rangle\langle n|\bar{Q}|m\rangle\langle m|\bar{N}|0\rangle
mncosh(Δmmn(tQt/2)),\displaystyle\sim\sum_{m\neq n}\cosh{\left(\Delta m_{mn}(t_{Q}-t/2)\right)}, (32)

where Δmn=EmEn\Delta_{mn}=E_{m}-E_{n}. Note that two states of nn and mm should have different intrinsic parities, otherwise the contribution vanishes since the matrix element should be zero, e.g., n|Q¯|n=0\langle n|\bar{Q}|n\rangle=0 in CP-even QCD vacuum if nn has even or odd parity. On the other hand, in the case of t<tQt<t_{Q}, it has

ΔC2pt(t,tQ)=Q¯(tQ)N(t)N¯(0)\displaystyle\Delta C_{\rm 2pt}(t,t_{Q})=\langle\bar{Q}(t_{Q})N(t)\bar{N}(0)\rangle n,meEntQEmt0|Q¯|nn|N|mm|N¯|0\displaystyle\sim\sum_{n,m}e^{-E_{n}t_{Q}-E_{m}t}\langle 0|\bar{Q}|n\rangle\langle n|N|m\rangle\langle m|\bar{N}|0\rangle
neEntQ,\displaystyle\sim\sum_{n^{\prime}}e^{-E_{n^{\prime}}t_{Q}}, (33)

where the state nn^{\prime} should be a P-odd state that couples to a nucleon state with a non-zero value of the matrix element 0|Q¯|n0\langle 0|\bar{Q}|n^{\prime}\rangle\neq 0. From the spectrum decomposition we see that the nucleon mixing angle α\alpha is a mixing parameter between the ground state nucleon and CP-odd excited states. To see the truncation artefacts, the authors of Refs. [15, 16] consider the partial summed two-point correlation function C2ptQ¯(ts)=tQ=tstsΔC2pt(t,tQ)\displaystyle C^{\bar{Q}}_{\rm 2pt}(t_{s})=\sum_{t_{Q}=-t_{s}}^{t_{s}}\Delta C_{\rm 2pt}(t,t_{Q}). From Eqs. (3) and (3) its asymptotic form behaves like C2ptQ¯(ts)=A+BeEts\displaystyle C^{\bar{Q}}_{\rm 2pt}(t_{s})=A+Be^{-Et_{s}} for tstt_{s}\gtrsim t, where the contribution for large tst_{s} is expected to be exponentially suppressed. This is numerically checked as shown in the left panel of Fig. 7, where a plateau is obtained for tstt_{s}\sim t and the contributions from ts>tt_{s}>t seem to be below the statistical fluctuation. Neglecting unnecessary noise from tstt_{s}\gtrsim t, an improvement of α\alpha up to a factor 2 is obtained. The same spectrum decomposition can be applied to the modified nucleon three point function ΔC3ptQ¯(t,tQ,top)=T{N(T)Jμ(top)Q¯(tQ)N¯(0)}\Delta C^{\bar{Q}}_{\rm 3pt}(t,t_{Q},t_{\rm op})=\langle{\rm T}\{N(T)J^{\mu}(t_{\rm op})\bar{Q}(t_{Q})\bar{N}(0)\}\rangle, where a fit analysis using its asymptotic form with the α\alpha-improvement yields a factor of 232\sim 3 times increases in the signal-to-noise ratio for EDFF F3(Q2)F_{3}(Q^{2}). The right panel of Fig. 7 shows a chiral (and continuum) extrapolation for dnθd_{n}^{\theta} using six ensembles with mπ>410m_{\pi}>410 MeV for several lattice spacings. A ChPT fit ansatz dnθmqmπ2d_{n}^{\theta}\sim m_{q}\sim m_{\pi}^{2} is used to further constrain the nEDM towards the chiral limit, which yields dnθ=0.00152(71)θd_{n}^{\theta}=-0.00152(71)\theta ee fm at the physical point with \sim 2 σ\sigma deviation from zero. This result is consistent with the aforementioned naively scaled value |dnθ(mπphys)|0.02e2mN|d_{n}^{\theta}(m_{\pi}^{\rm phys})|\lesssim 0.02\frac{e}{2m_{N}} [24]. Their fitted data, however, are in heavy pion mass region and do not clearly show the chiral behavior dnθmπ2d_{n}^{\theta}\propto m_{\pi}^{2}, so that the fit results seem to be less convincing. To avoid model dependence, more accurate results near the physical point are needed.

Refer to caption
Refer to caption
Figure 7: Results presented in Ref. [16]. (Left) The improved nucleon mixing angle α\alpha plotted against the sum parameter tst_{s}.  (Right) A chiral and continuum extraplation of nEDM dnθd_{n}^{\theta} plotted as a function of mπm_{\pi}. The bands are the fit results.

New results near the physical point (mπ=130m_{\pi}=130-310310 MeV) are presented in a parallel talk by Boram Yoon [17]. The topological charge is calculated using the O(a4)O(a^{4})-improved field strength [27] with the gradient flow on MILC HISQ ensembles with a=0.060.15a=0.06-0.15 fm. This study also uses the truncation method in tt-direction. While the convergence properties in partial sum can be seen in both two- and three- point functions, due to the slow convergence at the physical point no significant improvement is observed. It is, however, remarkable that the number of measurements is OO(100k), which gives a statistically significant signal for F3(Q2)F_{3}(Q^{2}) with non-zero Q2Q^{2} near the physical point (see the left panel of Fig. 8). It is also reported that a variance reduction technique introduced in [28] has about 2525% error reduction. In this calculation the excited state contamination is removed by using the two-state fits with multiple source-sink separations for each momentum Q2Q^{2}. The results for F3(Q2)F_{3}(Q^{2}) at heavier mπm_{\pi} with non-zero Q2Q^{2} are consistent with the previous lattice results [24, 16]. The result from the chiral (interpolation) and continuum extrapolation fit for dnθd_{n}^{\theta} are also presented in the right panel of Fig. 8, which yields a non-zero signal of |dnθ|=0.011(6)θ|d_{n}^{\theta}|=0.011(6)\theta ee fm at the physical point. While this result is also consistent with estimates from ChPT analysis and the QCD sum rules [4], it is not sufficient to constrain the θ\theta parameter. We also notice that even though the values of dnθd_{n}^{\theta} at finite aa are all positive except for the data at mπ=135m_{\pi}=135 MeV, these values in the continuum limit become negative, which may indicate a sizable discretization effect on dnθd_{n}^{\theta}. In addition, there is an increasing tendency of F3(Q2)F_{3}(Q^{2}) towards Q2,mπ20Q^{2},m_{\pi}^{2}\to 0 but with larger error. A leading order ChPT fit may not be suitable for dnθd_{n}^{\theta} in the simulation mass region, and understanding the Q2Q^{2} dependence of F3(Q2)F_{3}(Q^{2}) would be important to precisely determine dnθd_{n}^{\theta} at the physical point.

We have shown in this section that there are several ongoing studies of the θ\theta-induced nEDM. Even having OO(100k) statistics with employing noise reduction techniques at the physical point, it still is not sufficient to constrain the θ\theta parameter due to the large fluctuation of the topological charge density, which should become even worse when approaching to the continuum and larger volume limit. There also are several systematic uncertainties due to finite lattice spacing and Q20Q^{2}\to 0 extrapolation in EDFF F3(Q2)F_{3}(Q^{2}), which need to be further explored. Thus θ\theta-induced EDM at the physical point will be extremely challenging and will require more special techniques that work well in particular for gluon operators. In the next section, we would like to propose a new approach using a matrix element of the nucleon with background electric field. Since this approach is based on the energy shift method, we can directly obtain dnθd_{n}^{\theta} without Q2Q^{2} extrapolation, which may be potentially advantageous over the form factor method.

Refer to caption
Refer to caption
Figure 8: Preliminary results presented in Ref. [17]. (Left) The Q2Q^{2} dependence of F3(Q2)/(2mN)F_{3}(Q^{2})/(2m_{N}).  (Right) A chiral and continuum extraplation of nEDM dnθd_{n}^{\theta} plotted as a function of mπ2m_{\pi}^{2}.

4 New approach based on the matrix element with the background electric field

The idea is to simply apply the truncation technique to the energy shift method in the presence of the background electric field, in which we find that the energy shift δE\delta E in Eq. (25) is given as a nucleon matrix element of the topological charge density operator Q¯\bar{Q} in Eq. (30). Throughout the section, the state’s momenta are set to zero and arguments of p\vec{p} are suppressed. Performing the spectrum decomposition of ΔC2pt(t,tQ)\Delta C_{\rm 2pt}(t,t_{Q}) as in Eq. (3) but now in the presence of the background electric field, we obtain the following result for t>tQt>t_{Q} as

ΔC2pt,E(t,tQ)\displaystyle\Delta C_{{\rm 2pt},\vec{E}}(t,t_{Q}) =N(t)Q¯(tQ)N¯(0)E\displaystyle=\langle N(t)\bar{Q}(t_{Q})\bar{N}(0)\rangle_{\vec{E}}
=n,meEn(ttQ)EmtQ0|N|n,En,E|Q¯|m,Em,E|N¯|0\displaystyle=\sum_{n,m}e^{-E_{n}(t-t_{Q})-E_{m}t_{Q}}\langle 0|N|n,\vec{E}\rangle\langle n,\vec{E}|\bar{Q}|m,\vec{E}\rangle\langle m,\vec{E}|\bar{N}|0\rangle
|ZN+|2emN+tN+,E|Q¯|N+,E,\displaystyle\sim|Z_{N_{+}}|^{2}e^{-m_{N_{+}}t}\langle N_{+},\vec{E}|\bar{Q}|N_{+},\vec{E}\rangle, (34)

where |N+,E|N_{+},\vec{E}\rangle is the ground state nucleon in the presence of the background electric field. We should note that in contrast to Eq. (3), there is a leading order contribution from the ground state nucleon given as a matrix element N+,E|Q¯|N+,E\langle N_{+},\vec{E}|\bar{Q}|N_{+},\vec{E}\rangle. Again taking the partial summation of ΔC2pt,E(t,tQ)\Delta C_{{\rm 2pt},\vec{E}}(t,t_{Q}) over sink-source separation tt, we obtain

C2pt,EQ¯(t)=tQ=0tΔC2pt,E(t,tQ)|ZN+|2emN+t(tN+,E|Q¯|N+,E),\displaystyle C_{{\rm 2pt},\vec{E}}^{\bar{Q}}(t)=\sum_{t_{Q}=0}^{t}\Delta C_{{\rm 2pt},\vec{E}}(t,t_{Q})\sim|Z_{N_{+}}|^{2}e^{-m_{N_{+}}t}(t\langle N_{+},\vec{E}|\bar{Q}|N_{+},\vec{E}\rangle), (35)

where the matrix element is given the coefficient of linear in tt. Comparing Eq. (25) with Eq. (35), we find that the matrix element should correspond to the energy shift

N+,E|Q¯|N+,E=ζ2mNu¯[ΣE]u+O(E2).\displaystyle\langle N_{+},\vec{E}|\bar{Q}|N_{+},\vec{E}\rangle=-\frac{\zeta}{2m_{N}}\bar{u}\left[\vec{\Sigma}\cdot\vec{E}\right]u+O(\vec{E}^{2}). (36)

This formula is analogous to the leading order energy correction in the perturbation theory of quantum mechanics, c.f., ΔEn=n|ΔH^|n\Delta E_{n}=\langle n|\Delta\hat{H}|n\rangle for Hamiltonian H^=H^0+ΔH^\hat{H}=\hat{H}_{0}+\Delta\hat{H}. In this case we regard the electric field as a perturbation in addition to the CP-odd Hamiltonian (θ\theta-term), and also take into account a leading order perturbation effect on the state |N+,E|N_{+},\vec{E}\rangle. Even without θ\theta-term, the ground state |N+,E|N_{+},\vec{E}\rangle could mix with CP-odd states [29] due to the background electric field as

|N+,E\displaystyle|N_{+},\vec{E}\rangle =|N++cED|N+,\displaystyle=|N_{+}\rangle+c\vec{E}\cdot\vec{D}|N_{-}\rangle+\cdots, (37)

where |N+|N_{+}\rangle is the parity-even ground state nucleon in (lattice) CP-even vacuum. The leading order correction should come from a parity-odd nucleon |N|N_{-}\rangle with an overlap coefficients of cc and the electric field E\vec{E} and an expectation value of the dipole operator D\vec{D}  [29]. Substituting Eq. (37) into Eq. (36), we obtain

N+,E|Q¯|N+,E\displaystyle\langle N_{+},\vec{E}|\bar{Q}|N_{+},\vec{E}\rangle =EDN+|Q¯|N+(c.c.)+,\displaystyle=\vec{E}\cdot\vec{D}\langle N_{+}|\bar{Q}|N_{-}\rangle+({\rm c.c.})+\cdots, (38)

where we note N±|Q¯|N±=0\langle N_{\pm}|\bar{Q}|N_{\pm}\rangle=0 due to parity symmetry, since we consider the electric field as a perturbation so that |N±|N_{\pm}\rangle is defined in CP-even QCD vacuum. Thus the contribution of the matrix element N+|Q¯|N\langle N_{+}|\bar{Q}|N_{-}\rangle is exactly the same as the parity mixing effect that appears in calculation of α\alpha in Eq. (3). In the perturbation theory point of view, the EDM is an interplay of the electric field and the CP-odd operator both in the first order perturbation. Using the standard ratio method as in Eq. (26) and the relation to the matrix element in Eq. (36), we obtain the modified energy shift formula

RSz±(t,tQ)=Tr[TSz±ΔC2pt,E(t,tQ)]Tr[TpC2pt,E(t)]ζ2mNEz,(t),\displaystyle R_{S_{z\pm}}(t,t_{Q})=\frac{{\rm Tr}[T_{S_{z\pm}}\Delta C_{{\rm 2pt},\vec{E}}(t,t_{Q})]}{{\rm Tr}[T_{p}C_{2pt,\vec{E}}(t)]}\to\mp\frac{\zeta}{2m_{N}}E_{z},\quad\quad(t\to\infty), (39)

From this formula, it is clear that we do not need to extend tQt_{Q} outside the sink-source position, since there is no other term that is proportional to tt in tQ>tt_{Q}>t. In fact the contributions from tQ>tt_{Q}>t should be excited state contaminations that should disappear in the limit tt\to\infty. Thus, without Q2Q^{2} extrapolation, the EDFF F3(0)F_{3}(0) can be directly extracted from the ratio by dividing by the electric field as |F3(0)|=limt2mN|RSz±(t,tQ)||Ez|\displaystyle|F_{3}(0)|=\lim_{t\to\infty}\frac{2m_{N}|R_{S_{z\pm}}(t,t_{Q})|}{|E_{z}|}.

We show our preliminary results on the θ\theta-induced nEDM from the matrix element approach. Since we compute the matrix element with the electric fields along zz-direction in both positive and negative, we have four results for each component of ±Ez\pm E_{z} and projections TSz±T_{S_{z\pm}} (see Fig. 9). Obviously these data are correlated with each other and differences between spin up (or positive EzE_{z}) and spin down (or negative EzE_{z}) reduce the error. Fig. 10 shows the gradient flow time tgft_{gf} dependence of F3(0)F_{3}(0) for each sink-source time separation t=Tt=T. As expected, the signal becomes better as increasing the flow time, and the result for F3(0)F_{3}(0) becomes stable at tgf4t_{gf}\geq 4 and T8T\geq 8. We also study a possible higher order correction in E\vec{E}. Fig. 11 shows the comparison of two results for F3(0)F_{3}(0) with different electric field strength with |n|=1|n|=1 and 22, where both results |n|=1|n|=1 and 22 are consistent with each other for T8T\geq 8, This result indicates that O(E2)O(\vec{E}^{2}) corrections on F3(0)F_{3}(0) are small. From the plateau at T=9T=9 with tgf=8t_{gf}=8 we obtain F3(0)=0.12(3)F_{3}(0)=0.12(3) for the neutron at mπ=330m_{\pi}=330 MeV. Our result seems to be consistent with the previous result obtained in the form factor method (see e.g., Fig. 8) while our value is only available at Q2=0Q^{2}=0 and not directly comparable with the previous results.

Refer to caption
Refer to caption
Figure 9: Preliminary results of F3(0)F_{3}(0) from the matrix element approach. Results are obtained on a 243×6424^{3}\times 64 lattice with mπ=330m_{\pi}=330 MeV. 4 different components (Left) and their linear combinations for obtaining better signals (Right).
Refer to caption
Refer to caption
Figure 10: Gradient flow time dependence of F3(0)F_{3}(0) for each sink-source time separation TT. The horizontal axis represents tQT/2t_{Q}-T/2. Results are obtained on a 243×6424^{3}\times 64 lattice with mπ=330m_{\pi}=330 MeV and |n|=2|n|=2 (the electric field quanta).
Refer to caption
Refer to caption
Figure 11: Electric field dependence of F3(0)F_{3}(0) for |n|=1|n|=1 and 22 at a fixed gradient flow time. The horizontal axis represents tQT/2t_{Q}-T/2. Results are obtained on a 243×6424^{3}\times 64 lattice with mπ=430m_{\pi}=430 MeV.

5 Summary

Lattice calculations of nEDM are important for interpreting CP-violation effects in EDM experiments and cosmological observations. A number of groups are putting in the effort required for computing nEDM at the physical point using the form factor method. The nEDM induced by θ\theta-term has a large statistical noise in its correlation to the topological charge density, which is not suppressed at a large distance due to its global nature. To reduce the error several noise reduction techniques using the truncation of (space and) time region of the topological charge density have been employed. It is found that while the truncation reduces the error by a factor of 2 at a heavier pion mass, due to its poor convergence no significant improvement is observed at the physical pion mass. There also are several systematic uncertainties in F3(Q2)F_{3}(Q^{2}). The uncertainties of the discretization effect, Q20Q^{2}\to 0 extrapolation, and excited state contamination for F3(Q2)F_{3}(Q^{2}) have not been well understood, which seem to become significant near the physical point. To control the systematic errors and to further improve the statistical signal, we have proposed a new approach using the matrix element with background electric fields. This method only requires a local topological charge density operator between sink and source positions, and thus can avoid the large topological noise at a large distance. In addition, no Q2Q^{2} extrapolation is required since the forward matrix element is directly obtained from the energy shift. Our preliminary results have demonstrated that we can achieve statistically-significant signal at heavier pion masses that are consistent with the previous results. This method can in principle be applied to any CP{\cancel{\rm CP}} operators, in which the Weinberg’s three-gluon operator especially is beneficial for this method, since there is no additional computation cost for the gluonic operator. We need further investigations at the physical point.

Acknowledgments

The speaker would like to thank the organizers of Lattice2019 in Wuhan for the invitation. We would also like to thank Tanmoy Bhattacharya and Boram Yoon for sending us their materials and the fruitful discussions. We are grateful for the gauge configurations provided by the RBC/UKQCD collaboration. This research used resources of the Argonne Leadership Computing Facility, which is a DOE Office of Science User Facility supported under Contract DE-AC02-06CH11357, and Hokusai supercomputer of the RIKEN ACCC facility. H.O. is supported in part by JSPS KAKENHI Grant Numbers 17K14309 and 18H03710. S.S. is supported by the National Science Foundation under CAREER Award PHY-1847893.

References

  • [1] C. A. Baker et al., “An Improved experimental limit on the electric dipole moment of the neutron,” Phys. Rev. Lett.  97, 131801 (2006) doi:10.1103/PhysRevLett.97.131801 [hep-ex/0602020].
  • [2] B. Graner, Y. Chen, E. G. Lindahl and B. R. Heckel, “Reduced Limit on the Permanent Electric Dipole Moment of Hg199,” Phys. Rev. Lett.  116, no. 16, 161601 (2016) Erratum: [Phys. Rev. Lett.  119, no. 11, 119901 (2017)] doi:10.1103/PhysRevLett.119.119901, 10.1103/PhysRevLett.116.161601 [arXiv:1601.04339 [physics.atom-ph]].
  • [3] Y. Nakamura and G. Schierholz, arXiv:1912.03941 [hep-lat].
  • [4] J. Engel, M. J. Ramsey-Musolf and U. van Kolck, “Electric Dipole Moments of Nucleons, Nuclei, and Atoms: The Standard Model and Beyond,” Prog. Part. Nucl. Phys.  71, 21 (2013) doi:10.1016/j.ppnp.2013.03.003 [arXiv:1303.2371 [nucl-th]].
  • [5] N. Yamanaka, “Review of the electric dipole moment of light nuclei,” Int. J. Mod. Phys. E 26, no. 4, 1730002 (2017) doi:10.1142/S0218301317300028 [arXiv:1609.04759 [nucl-th]].
  • [6] T. Bhattacharya, R. Gupta, B. Yoon, ”Recent results of nucleon strucutre & matrix element calculations” Proceedings of Sciece, LATTICE2019 (2020), p.247.
  • [7] R. Gupta, “Present and future prospects for lattice QCD calculations of matrix elements for nEDM,” PoS SPIN 2018, 095 (2019) doi:10.22323/1.346.0095 [arXiv:1904.00323 [hep-lat]].
  • [8] E. Shintani et al., “Neutron electric dipole moment from lattice QCD,” Phys. Rev. D 72, 014504 (2005) doi:10.1103/PhysRevD.72.014504 [hep-lat/0505022].
  • [9] F. Berruto, T. Blum, K. Orginos and A. Soni, “Calculation of the neutron electric dipole moment with two dynamical flavors of domain wall fermions,” Phys. Rev. D 73, 054509 (2006) doi:10.1103/PhysRevD.73.054509 [hep-lat/0512004].
  • [10] S. Aoki et al., “The Electric dipole moment of the nucleon from simulations at imaginary vacuum angle theta,” arXiv:0808.1428 [hep-lat].
  • [11] F.-K. Guo et al., “The electric dipole moment of the neutron from 2+1 flavor lattice QCD,” Phys. Rev. Lett.  115, no. 6, 062001 (2015) doi:10.1103/PhysRevLett.115.062001 [arXiv:1502.02295 [hep-lat]].
  • [12] C. Alexandrou et al., “Neutron electric dipole moment using Nf=2+1+1N_{f}=2+1+1 twisted mass fermions,” Phys. Rev. D 93, no. 7, 074503 (2016) doi:10.1103/PhysRevD.93.074503 [arXiv:1510.05823 [hep-lat]].
  • [13] E. Shintani, T. Blum, T. Izubuchi and A. Soni, “Neutron and proton electric dipole moments from Nf=2+1N_{f}=2+1 domain-wall fermion lattice QCD,” Phys. Rev. D 93, no. 9, 094503 (2016) doi:10.1103/PhysRevD.93.094503 [arXiv:1512.00566 [hep-lat]].
  • [14] M. Abramczyk, S. Aoki, T. Blum, T. Izubuchi, H. Ohki and S. Syritsyn, “Lattice calculation of electric dipole moments and form factors of the nucleon,” Phys. Rev. D 96, no. 1, 014501 (2017) doi:10.1103/PhysRevD.96.014501 [arXiv:1701.07792 [hep-lat]].
  • [15] J. Dragos, T. Luu, A. Shindler, J. de Vries and A. Yousif, “Improvements to Nucleon Matrix Elements within a θ\theta Vacuum from Lattice QCD,” PoS LATTICE 2018, 259 (2019) doi:10.22323/1.334.0259 [arXiv:1809.03487 [hep-lat]].
  • [16] J. Dragos, T. Luu, A. Shindler, J. de Vries and A. Yousif, “Confirming the Existence of the strong CP Problem in Lattice QCD with the Gradient Flow,” arXiv:1902.03254 [hep-lat].
  • [17] B. Yoon, T. Bhattacharya, V. Cirigliano and R. Gupta, “Neutron Electric Dipole Moments with Clover Fermions,” PoS LATTICE 2019, 243 (2019) [arXiv:2003.05390 [hep-lat]].
  • [18] E. Shintani et al., “Neutron electric dipole moment with external electric field method in lattice QCD,” Phys. Rev. D 75, 034507 (2007) doi:10.1103/PhysRevD.75.034507 [hep-lat/0611032].
  • [19] E. Shintani, S. Aoki and Y. Kuramashi, “Full QCD calculation of neutron electric dipole moment with the external electric field method,” Phys. Rev. D 78, 014503 (2008) doi:10.1103/PhysRevD.78.014503 [arXiv:0803.0797 [hep-lat]].
  • [20] T. Izubuchi, S. Aoki, K. Hashimoto, Y. Nakamura, T. Sekido and G. Schierholz, “Dynamical QCD simulation with theta terms,” PoS LATTICE 2007, 106 (2007) doi:10.22323/1.042.0106 [arXiv:0802.1470 [hep-lat]].
  • [21] W. Detmold, B. C. Tiburzi and A. Walker-Loud, “Extracting Electric Polarizabilities from Lattice QCD,” Phys. Rev. D 79, 094505 (2009) doi:10.1103/PhysRevD.79.094505 [arXiv:0904.1586 [hep-lat]].
  • [22] W. Detmold, B. C. Tiburzi and A. Walker-Loud, “Extracting Nucleon Magnetic Moments and Electric Polarizabilities from Lattice QCD in Background Electric Fields,” Phys. Rev. D 81, 054502 (2010) doi:10.1103/PhysRevD.81.054502 [arXiv:1001.1131 [hep-lat]].
  • [23] K. F. Liu, J. Liang and Y. B. Yang, “Variance Reduction and Cluster Decomposition,” Phys. Rev. D 97, no. 3, 034507 (2018) doi:10.1103/PhysRevD.97.034507 [arXiv:1705.06358 [hep-lat]].
  • [24] S. Syritsyn, T. Izubuchi and H. Ohki, “Calculation of Nucleon Electric Dipole Moments Induced by Quark Chromo-Electric Dipole Moments and the QCD θ\theta-term,” PoS Confinement 2018, 194 (2019) doi:10.22323/1.336.0194 [arXiv:1901.05455 [hep-lat]].
  • [25] T. Blum, T. Izubuchi and E. Shintani, “New class of variance-reduction techniques using lattice symmetries,” Phys. Rev. D 88, no. 9, 094503 (2013) doi:10.1103/PhysRevD.88.094503 [arXiv:1208.4349 [hep-lat]].
  • [26] P. de Forcrand, M. Garcia Perez and I. O. Stamatescu, “Topology of the SU(2) vacuum: A Lattice study using improved cooling,” Nucl. Phys. B 499, 409 (1997) doi:10.1016/S0550-3213(97)00275-7 [hep-lat/9701012].
  • [27] S. O. Bilson-Thompson, D. B. Leinweber and A. G. Williams, “Highly improved lattice field strength tensor,” Annals Phys.  304, 1 (2003) doi:10.1016/S0003-4916(03)00009-5 [hep-lat/0203008].
  • [28] T. Bhattacharya, B. Yoon, R. Gupta and V. Cirigliano, “Neutron Electric Dipole Moment from Beyond the Standard Model,” arXiv:1812.06233 [hep-lat].
  • [29] G. Baym and D. H. Beck, “Elementary quantum mechanics of the neutron with an electric dipole moment,” Proc. Nat. Acad. Sci.  113, 7438 (2016) doi:10.1073/pnas.1607599113 [arXiv:1606.06774 [hep-ph]].