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

Constraints on neutrino non-standard interactions from COHERENT and PandaX-4T

Gang Li [email protected] School of Physics and Astronomy, Sun Yat-sen University, Zhuhai 519082, P.R. China    Chuan-Qiang Song [email protected] School of Fundamental Physics and Mathematical Sciences, Hangzhou Institute for Advanced Study, UCAS, Hangzhou 310024, China School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, P.R. China Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, P. R. China    Feng-Jie Tang [email protected] School of Fundamental Physics and Mathematical Sciences, Hangzhou Institute for Advanced Study, UCAS, Hangzhou 310024, China School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, P.R. China Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, P. R. China    Jiang-Hao Yu [email protected] School of Fundamental Physics and Mathematical Sciences, Hangzhou Institute for Advanced Study, UCAS, Hangzhou 310024, China School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, P.R. China Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, P. R. China International Centre for Theoretical Physics Asia-Pacific, Beijing/Hangzhou, China
Abstract

We investigate constraints on neutrino non-standard interactions (NSIs) in the effective field theory framework, using data from the first measurement of solar 8B neutrinos via coherent elastic neutrino-nucleus scattering (CEν\nuNS) in the PandaX-4T experiment and the COHERENT experiment. In the PandaX-4T experiment, due to relatively large statistical uncertainties and measured CEν\nuNS counts that significantly differ from the Standard Model predictions, its sensitivities to the neutrino NSIs are currently limited, compared to the COHERENT experiment. However, the PandaX-4T experiment is uniquely sensitive to the neutrino NSIs for the τ\tau flavor due to oscillation feature of the solar 8B neutrinos. We also assess how the experimental central value, exposure, and systematic uncertainties will affect the constraints on neutrino NSIs from various CEν\nuNS measurements in the future.

I Introduction

In the Standard Model (SM), neutrinos interact with ordinary matter through the exchange of WW and ZZ bosons. Neutrino non-standard interactions (NSIs) in the charged and neutral currents beyond the SM (BSM) were initially formulated by Lee-Yang Lee and Yang (1956) and Wolfenstein Wolfenstein (1978), respectively. While the detection of neutrinos in neutrino oscillation experiments only involves charged-current interactions, the coherent elastic neutrino-nucleus scattering (CEν\nuNS) Freedman (1974) serves as a unique probe of the neutral-current neutrino NSIs, potentially arising from new mediators, such as ZZ^{\prime} boson Barranco et al. (2005).

Despite the coherent enhancement of the CEν\nuNS cross section, this process is difficult to detect due to small deposited energy. It was first observed in the COHERENT experiment using the CsI[Na] scintillation detector Akimov et al. (2017) and later argon Akimov et al. (2021) and germanium Adamski et al. (2024) detectors with neutrinos produced from the spallation neutron source (SNS). These results as well as the follow-up detection with a larger exposure of CsI[Na] Akimov et al. (2022a) and other experimental efforts Aguilar-Arevalo et al. (2019); Bonet et al. (2021); Alekseev et al. (2022); Colaresi et al. (2022); Su et al. (2023); Augier et al. (2023); Ackermann et al. (2024); Yang et al. (2024); Cai et al. (2024); Xiao (2024) have motivated diverse phenomenological studies Coloma et al. (2017a); Liao and Marfatia (2017); Cadeddu et al. (2018); Ge and Shoemaker (2018); Aristizabal Sierra et al. (2018a); Ciuffoli et al. (2018); Farzan et al. (2018); Billard et al. (2018); Aristizabal Sierra et al. (2018b); Cadeddu and Dordei (2019); Altmannshofer et al. (2019); Miranda et al. (2019); Aristizabal Sierra et al. (2019); Papoulias (2020); Giunti (2020); Canas et al. (2020); Hoferichter et al. (2020); Skiba and Xia (2022); Cadeddu et al. (2021); Du et al. (2022); Dasgupta et al. (2021); Atzori Corona et al. (2022); De Romeri et al. (2023); Li et al. (2024), see also Refs. Lindner et al. (2017); Dent et al. (2017); Coloma et al. (2017b) for earlier studies and Ref. Abdullah et al. (2022) for a recent review.

On the other hand, with the tremendous progress in the sensitivity of dark matter (DM) direct detection, it is anticipated that the DM experiments can detect neutrinos from astrophysical sources. These neutrinos exhibit nuclear recoil signatures resembling those of DM, which pose as irreducible backgrounds in DM direct detection and are often referred to as the “neutrino frog” Billard et al. (2014); O’Hare (2021); Tang and Zhang (2023). This is not unexpected since the idea of detecting DM that scatters off nuclei Goodman and Witten (1985) was inspired by the proposal to detect MeV-range neutrinos via CEν\nuNS Drukier and Stodolsky (1984).

Recently, the solar 8B neutrino was measured through CEν\nuNS in the PandaX-4T Bo et al. (2024) and XENONnT Aprile et al. (2024) experiments with corresponding statistical significance of 2.64σ2.64\sigma and 2.73σ2.73\sigma, which signify the first step into the neutrino frog experimentally. Assuming the SM cross section of CEν\nuNS, the signals are interpreted as the measurements of solar 8B neutrino flux of (8.4±3.1)×106cm2s1(8.4\pm 3.1)\times 10^{6}\mathrm{\leavevmode\nobreak\ cm}^{-2}\mathrm{\leavevmode\nobreak\ s}^{-1} and (4.72.3+3.6)×106cm2s1\left(4.7_{-2.3}^{+3.6}\right)\times 10^{6}\mathrm{\leavevmode\nobreak\ cm}^{-2}\mathrm{\leavevmode\nobreak\ s}^{-1}, respectively, both of which are consistent with the standard solar model predictions Vinyoles et al. (2017); Bahcall et al. (1996, 2006) and the results from dedicated solar neutrino experiments Aharmim et al. (2013); Abe et al. (2011, 2016); Agostini et al. (2020).

In the SM, empirical nuclear form factors Helm (1956); Klein and Nystrand (1999) that parameterize the nuclear response are usually adopted to calculate the cross section of CEν\nuNS. However, an improved treatment is necessary if neutrino NSIs are present, which is feasible in the effective field theory (EFT) approach Altmannshofer et al. (2019); Skiba and Xia (2022); Hoferichter et al. (2020)111For vector and axial-vector NSIs, one can also refine the CEν\nuNS cross section by modifying the weak charge Barranco et al. (2005); Abdullah et al. (2022). , analogous to the situation of DM-nucleus scattering Cirigliano et al. (2012); Menendez et al. (2012); Klos et al. (2013); Vietze et al. (2015); Hoferichter et al. (2015); Fitzpatrick et al. (2013); Anand et al. (2014); Bishara et al. (2017a, b).

An end-to-end EFT framework Altmannshofer et al. (2019) was developed and utilized to describe the CEν\nuNS process from the new physics scale to the nuclear scale, which takes advantages of the heavy baryon chiral perturbation theory (HBChPT) Jenkins and Manohar (1991) and multipole expansions for nuclear responses Walecka (1995)222Indeed, the multipole analysis was applied to neutrino-nucleus scattering processes in the SM weak charged currents slightly before the proposal of CEν\nuNS O’ Connell et al. (1972). The effect of SM weak neutral currents in neutrino scattering off nuclei was later investigated in Ref. Donnelly and Peccei (1979). . Compared to the nuclear form factor approach, the EFTs enable controlled uncertainties in systematic power countings, which allows for potential theoretical improvements in nuclear shell-model calculations Hoferichter et al. (2020) and beyond Abdel Khaleq et al. (2024).

In this work, we will investigate the sensitivities to neutrino NSIs in the EFT framework, including matching between several EFTs. We consider relevant dimension-6 operators in the low-energy effective field (LEFT) and QCD chiral Lagrangian with external sources and heavy baryon expansion, and finally match to nuclear response function to obtain the most stringent constraints from the measurement conducted with the CsI[Na] detector in the COHERENT experiment Akimov et al. (2022a) and the first result of new physics using the measurement of solar 8B neutrinos by PandaX-4T Bo et al. (2024).

The remainder of the paper is organized as follows. In Sec. II, we discuss the neutrino NSIs from quark level to the nucleon level using the LEFT and HBChPT. In Sec. III, we derive the CEν\nuNS cross section using the multipole expansions for nuclear responses. In Sec. IV, we evaluate the event rates of the CEν\nuNS signals in the COHERENT and PandaX-4T experiments (considering the matter effects in the neutrino propagation by using package PEANUTS), and obtain the constraints on the Wilson coefficients for specific neutrino flavors, which are interpreted as lower bounds on the NSI energy scale using the χ2\chi^{2} analysis. Two-dimensional constraint on the NSI parameters is also obtained for the comparison of these two CEν\nuNS experiments. We have an assessment of the sensitivities of future measurements of solar 8B neutrinos via CEν\nuNS in DM detectors. We conclude in Sec. V. In the appendix, we provide details of the detector resolution and efficiency in COHERENT CsI measurement.

II Neutrino non-standard interactions

The neutral-current (NC) neutrino-quark interactions can be parameterized as Wolfenstein (1978); Scholberg (2006); Barranco et al. (2005); Davidson et al. (2003); Du et al. (2022)

NC\displaystyle\mathcal{L}_{\mathrm{NC}} 22GF[ϵαβqL(ν¯αγμPLνβ)(q¯γμPLq)\displaystyle\supset-2\sqrt{2}G_{F}\left[\epsilon_{\alpha\beta}^{qL}\left(\bar{\nu}_{\alpha}\gamma^{\mu}P_{L}\nu_{\beta}\right)\left(\bar{q}\gamma_{\mu}P_{L}q\right)\right.
+ϵαβqR(ν¯αγμPLνβ)(q¯γμPRq)],\displaystyle\quad\left.+\epsilon_{\alpha\beta}^{qR}\left(\bar{\nu}_{\alpha}\gamma^{\mu}P_{L}\nu_{\beta}\right)\left(\bar{q}\gamma_{\mu}P_{R}q\right)\right]\;, (1)

where PL/R=(1γ5)/2P_{L/R}=(1\mp\gamma_{5})/2, α,β\alpha,\beta denote the flavors of neutrinos, and q=u,dq=u,d, GFG_{F} is the Fermi constant.

In the LEFT, the relevant effective Lagrangian is

LEFT𝒞^1,q(6)𝒪1,q(6)+𝒞^2,q(6)𝒪2,q(6),\mathcal{L}_{\rm LEFT}\supset\hat{\cal C}_{1,q}^{(6)}{\cal O}_{1,q}^{(6)}+\hat{\cal C}_{2,q}^{(6)}{\cal O}_{2,q}^{(6)}\,, (2)

where the dimension-6 operators are defined as Jenkins et al. (2018); Altmannshofer et al. (2019)

𝒪1,q(6)\displaystyle{\cal O}_{1,q}^{(6)} =(ν¯αγμPLνβ)(q¯γμq),\displaystyle=(\bar{\nu}_{\alpha}\gamma_{\mu}P_{L}\nu_{\beta})(\bar{q}\gamma^{\mu}q),
𝒪2,q(6)\displaystyle{\cal O}_{2,q}^{(6)} =(ν¯αγμPLνβ)(q¯γμγ5q).\displaystyle=(\bar{\nu}_{\alpha}\gamma_{\mu}P_{L}\nu_{\beta})(\bar{q}\gamma^{\mu}\gamma_{5}q)\,. (3)

The correspondence between the coefficients ϵαβqL(R)\epsilon_{\alpha\beta}^{qL(R)} and the Wilson coefficients 𝒞^1,q(6)\hat{\cal C}_{1,q}^{(6)} and 𝒞^2,q(6)\hat{\cal C}_{2,q}^{(6)} is

ϵαβqL/R\displaystyle\epsilon_{\alpha\beta}^{qL/R} =122GF(𝒞^1,q(6)𝒞^2,q(6)).\displaystyle=-\dfrac{1}{2\sqrt{2}G_{F}}\left(\hat{\mathcal{C}}_{1,q}^{(6)}\mp\hat{\mathcal{C}}_{2,q}^{(6)}\right)\;. (4)

The SM contributions to the Wilson coefficients after integrating out the ZZ boson are Altmannshofer et al. (2019); Abdullah et al. (2022)

𝒞^1,u(d)(6)|SM\displaystyle\left.\hat{\mathcal{C}}_{1,u(d)}^{(6)}\right|_{\mathrm{SM}} =GF2(18(4)3sin2θW)δαβ,\displaystyle=\mp\frac{G_{F}}{\sqrt{2}}\left(1-\frac{8(4)}{3}\sin^{2}\theta_{W}\right)\delta_{\alpha\beta}\;, (5)
𝒞^2,u(d)(6)|SM\displaystyle\left.\hat{\mathcal{C}}_{2,u(d)}^{(6)}\right|_{\mathrm{SM}} =±GF2δαβ.\displaystyle=\pm\frac{G_{F}}{\sqrt{2}}\delta_{\alpha\beta}\;. (6)

where θW\theta_{W} is the weak mixing angle with sin2θW=0.2312\sin^{2}\theta_{W}=0.2312 Workman et al. (2022). We consider contributions from neutrino NSIs, so that

𝒞^i(6)=𝒞^i(6)|SM+𝒞^i(6)|NSI.\displaystyle\hat{\mathcal{C}}_{i}^{(6)}=\left.\hat{\mathcal{C}}_{i}^{(6)}\right|_{\rm SM}+\left.\hat{\mathcal{C}}_{i}^{(6)}\right|_{\rm NSI}\;. (7)

The dimensionful Wilson coefficient can also be expressed as

𝒞^i(6)|NSI=1ΛNSI2𝒞i(6)|NSI,\displaystyle\left.\hat{\mathcal{C}}_{i}^{(6)}\right|_{\rm NSI}=\dfrac{1}{\Lambda_{\rm NSI}^{2}}\left.\mathcal{C}_{i}^{(6)}\right|_{\rm NSI}\;, (8)

where 𝒞i(6)|NSI\left.\mathcal{C}_{i}^{(6)}\right|_{\rm NSI} is dimensionless, and the energy scale ΛNSI\Lambda_{\rm NSI} is determined by the mass of mediator that is responsible for the neutrino NSIs333For light mediator, there is additional momentum dependence from its propagator..

To obtain the neutrino-nucleus cross section of CEν\nuNS, we need to consider the matching in two steps Altmannshofer et al. (2019); Hoferichter et al. (2020): (1) from the quark level to the nucleon level; (2) from the nucleon level to the nucleus level. In the first step, the nucleons in the target are considered non-relativistic due to the small momentum exchange qq compared to the nucleon mass. The interaction Lagrangian for the neutrinos and non-relativistic nucleons is

NR=ci,N(d)𝒪i,N(d),\displaystyle\mathcal{L}_{\rm NR}=c_{i,N}^{(d)}\mathcal{O}_{i,N}^{(d)}\;, (9)

where N=p,nN=p,n, and dd denotes the number of derivatives in the operator.

By using the HBChPT, the following neutrino-nucleon operators at leading order are obtained:

𝒪1,N(0)\displaystyle\mathcal{O}_{1,N}^{(0)} =(ν¯βγμPLνα)(vμN¯vNv),\displaystyle=(\bar{\nu}_{\beta}\gamma_{\mu}P_{L}\nu_{\alpha})(v^{\mu}\bar{N}_{v}N_{v})\,, (10)
𝒪2,N(0)\displaystyle\mathcal{O}_{2,N}^{(0)} =(ν¯βγμPLνα)(N¯vSμNv).\displaystyle=(\bar{\nu}_{\beta}\gamma_{\mu}P_{L}\nu_{\alpha})(\bar{N}_{v}S^{\mu}N_{v})\,. (11)

where NvN_{v} denotes the large component of the nucleon field, and vμv^{\mu} is the nucleon velocity, the spin operator Sμ=γ5γμ/2S^{\mu}=\gamma_{5}\gamma_{\perp}^{\mu}/2 with γμγμvμ/v\gamma_{\perp}^{\mu}\equiv\gamma^{\mu}-v^{\mu}/v. In the lab frame, vμ=(1,0)v^{\mu}=(1,\vec{0}), and Sμ=(0,σ/2)S^{\mu}=(0,\vec{\sigma}/2) with σ\vec{\sigma} being the Pauli matrices. Denoting p1p_{1} and p2p_{2} (k1k_{1} and k2k_{2}) as the momenta of income and outcome neutrinos (nucleons), respectively, we can define the momentum transfer as q=p1p2q=p_{1}-p_{2}.

From the quark-level interactions to the nucleon-level interactions, the matching conditions are expressed as444Note that we use the symbol qq to represent both the momentum transfer and the quark in the conventional manner.

c1,N(0)\displaystyle c_{1,N}^{(0)} =q=u,dF1q/N𝒞^1,q(6),\displaystyle=\sum_{q=u,d}F_{1}^{q/N}\hat{\cal C}_{1,q}^{(6)}\,, c2,N(0)\displaystyle c_{2,N}^{(0)} =2q=u,dFAq/N𝒞^2,q(6),\displaystyle=2\sum_{q=u,d}F_{A}^{q/N}\hat{\cal C}_{2,q}^{(6)}\,, (12)

where FiF_{i} denote the momentum-dependent nucleon form factors describing the hadronization of quark currents. We use the values of FiF_{i} evaluated at q20q^{2}\to 0 Bishara et al. (2017b), which are accurate enough for our purpose,

F1u/p\displaystyle F_{1}^{u/p} =2,\displaystyle=2\;, F1d/p\displaystyle F_{1}^{d/p} =1,\displaystyle=1\;,
FAu/p\displaystyle F_{A}^{u/p} =0.897,\displaystyle=0.897\;, FAd/p\displaystyle F_{A}^{d/p} =0.376.\displaystyle=-0.376\;. (13)

III Cross section of CEν\nuNS

In the second step, the nuclear response to neutrino scattering needs to be considered at the nuclear level, which is described similarly to DM detection Fitzpatrick et al. (2013); Anand et al. (2014). In this framework, the many-body nuclear matrix elements are expanded using the multipole expansions Walecka (1995) in the harmonic oscillator basis, and can be calculated in the nuclear shell model Haxton and Lunardini (2008); Hoferichter et al. (2020).

To this end, we classify the Lagrangian terms according to charge operator (1N1_{N}) and nuclear spin operator (S\vec{S}Altmannshofer et al. (2019):

NR\displaystyle\mathcal{L}_{\rm NR} =(ν¯αl0,NPLνβ)1N+(ν¯αl5,NPLνβ)(2S),\displaystyle=\big{(}\bar{\nu}_{\alpha}l_{0,N}P_{L}\nu_{\beta}\big{)}1_{N}+\big{(}\bar{\nu}_{\alpha}\vec{l}_{5,N}P_{L}\nu_{\beta}\big{)}\cdot(2\vec{S})\,, (14)

where the Dirac structures are given by

l0,N=c1,N(0)/v,l5,Nμ=12c2,N(0)γμ,\displaystyle l_{0,N}=c_{1,N}^{(0)}/{v}\,,\quad l_{5,N}^{\mu}=\frac{1}{2}c_{2,N}^{(0)}\gamma^{\mu}\;, (15)

and l5\vec{l}_{5} in Eq. (14) is the spatial three-vector components of l5μl_{5}^{\mu}.

The differential cross section in the rest frame of the target nucleus is Altmannshofer et al. (2019)555We have corrected a missing factor of 1/81/8 in Eq.(3.23) of Ref. Altmannshofer et al. (2019), and have verified it by comparing the SM result with the calculation using the nuclear form factor Lindner et al. (2017); Abdullah et al. (2022).

dσdTnr=M8πEν2|¯|RW2,\frac{\rm{d}\sigma}{{\rm d}T_{\rm nr}}=\frac{M}{8\pi E_{\nu}^{2}}|\overline{{\cal M}}|_{RW}^{2}\;, (16)

where MM is the target nucleus mass, EνE_{\nu} is the initial neutrino energy, TnrT_{\rm nr} is the nuclear recoil energy, and the spin-averaged amplitude square is expressed as

|¯|RW2=4π2JA+1τ,τ=0,1(RMττWMττ\displaystyle|\overline{{\cal M}}|_{RW}^{2}=\frac{4\pi}{2J_{A}+1}\sum_{\tau,\tau^{\prime}=0,1}\Big{(}R_{M}^{\tau\tau^{\prime}}W_{M}^{\tau\tau^{\prime}} (17)
+RΣ′′ττWΣ′′ττ+RΣττWΣττ).\displaystyle+R_{\Sigma^{\prime\prime}}^{\tau\tau^{\prime}}W_{\Sigma^{\prime\prime}}^{\tau\tau^{\prime}}+R_{\Sigma^{\prime}}^{\tau\tau^{\prime}}W_{\Sigma^{\prime}}^{\tau\tau^{\prime}}\Big{)}\;.

Here, JAJ_{A} is the spin of the target nucleus, WiW_{i} denotes the nucleus response functions Fitzpatrick et al. (2013); Anand et al. (2014), and the kinematic factors RiR_{i} are given by Altmannshofer et al. (2019)

RMττ=Tr(PL/p1γ0l0,τγ0/p2l0,τ),\displaystyle\begin{split}R_{M}^{\tau\tau^{\prime}}&=\mathrm{Tr}\big{(}P_{L}/p_{1}\gamma_{0}l_{0,\tau^{\prime}}^{\dagger}\gamma_{0}/p_{2}l_{0,\tau}\big{)}\,,\end{split} (18)
RΣ′′ττ=Tr(PL/p1γ0l5,τjγ0/p2l5,τi)q^iq^j,\displaystyle\begin{split}R_{\Sigma^{\prime\prime}}^{\tau\tau^{\prime}}&=\mathrm{Tr}\big{(}P_{L}/p_{1}\gamma_{0}l_{5,\tau^{\prime}}^{j\dagger}\gamma_{0}/p_{2}l_{5,\tau}^{i}\big{)}\hat{q}^{i}\,\hat{q}^{j}\,,\end{split} (19)
RΣττ\displaystyle R_{\Sigma^{\prime}}^{\tau\tau^{\prime}} =Tr(PL/p1γ0l5,τjγ0/p2l5,τi)(δijq^iq^j),\displaystyle=\mathrm{Tr}\big{(}P_{L}/{p}_{1}\gamma_{0}l_{5,\tau^{\prime}}^{j\dagger}\gamma_{0}/p_{2}l_{5,\tau}^{i}\big{)}\big{(}\delta^{ij}-\hat{q}^{i}\,\hat{q}^{j}\big{)}\,, (20)

where τ,τ\tau,\tau^{\prime} are the isospin indices, q^q/|q|\hat{q}\equiv\vec{q}/|\vec{q}|, and i,j=1,2,3i,j=1,2,3. In the isospin basis, l0,τ=[l0,p+(1)τl0,n]/2l_{0,\tau}=\left[l_{0,p}+(-1)^{\tau}l_{0,n}\right]/2 and l5,τμ=[l5,pμ+(1)τl5,nμ]/2l_{5,\tau}^{\mu}=\left[l_{5,p}^{\mu}+(-1)^{\tau}l_{5,n}^{\mu}\right]/2.

IV Event rates and constraints

In the CEν\nuNS experiments, neutrinos from the source will interact with detector target nuclei, causing nucleus recoils. The resulting signal can be translated into the event rate. In the following, we will investigate constraints on neutrino NSIs from the measurements by the COHERENT and PandaX-4T experiments.

IV.1 Constraints from COHERENT

We first consider the measurements of the CEν\nuNS process in the COHERENT experiment using CsI[Na] Akimov et al. (2022a)666We do not consider the COHERENT measurement using Ar detector Akimov et al. (2021) since its sensitivity cannot compete with that CsI detector, even though the combination of these measurements can break degeneracy between different NSI parameter combinations De Romeri et al. (2023). . The differential event rate per target for neutrino flavor να=νe,νμ,ν¯μ\nu_{\alpha}=\nu_{e},\nu_{\mu},\bar{\nu}_{\mu} is expressed as Abdel Khaleq et al. (2024); Altmannshofer et al. (2019):

dRναdTnr=Eν,minEν,max𝑑EνΦνα(Eν)dσdTnr,\frac{{{\rm d}}R_{\nu_{\alpha}}}{{{\rm d}}T_{\rm nr}}=\int_{E_{\nu,{\rm min}}}^{E_{\nu,{\rm max}}}dE_{\nu}\Phi_{\nu_{\alpha}}(E_{\nu})\frac{{\rm d}\sigma}{{\rm d}T_{\rm nr}}\,, (21)

where dσ/dTnr{{\rm d}\sigma}/{{\rm d}T_{\rm nr}} is the differential cross section given in Eq. (16). The minimum initial neutrino energy is Eν,minMTnr/2E_{\nu,{\rm min}}\simeq\sqrt{MT_{\rm nr}/2}, where MM is the nucleus mass. The upper integration limit is given by the maximal energy of initial neutrinos produced in π+νμ(μ+e+νeν¯μ)\pi^{+}\rightarrow\nu_{\mu}(\mu^{+}\rightarrow e^{+}\nu_{e}\bar{\nu}_{\mu}). For νe\nu_{e} and ν¯μ\bar{\nu}_{\mu}, Eν,max=mμ/252.8MeVE_{\nu,{\rm max}}=m_{\mu}/2\simeq 52.8\leavevmode\nobreak\ {\rm{MeV}}, while for νμ\nu_{\mu}, Eν,max=(mπ2mμ2)/(2mπ)30MeVE_{\nu,{\rm max}}=(m_{\pi}^{2}-m_{\mu}^{2})/(2m_{\pi})\simeq 30\leavevmode\nobreak\ {\rm{MeV}} Atzori Corona et al. (2022); Aristizabal Sierra et al. (2018b), where mμm_{\mu} and mπm_{\pi} are the mass of the muon and pion, respectively.

The total neutrino fluxes are described by the Michel spectrum Coloma et al. (2017b); Liao and Marfatia (2017)

Φνe(Eν)\displaystyle\Phi_{\nu_{e}}(E_{\nu}) =𝒩192Eν2mμ3(12Eνmμ),\displaystyle={\cal N}\frac{192E_{\nu}^{2}}{m_{\mu}^{3}}\left(\frac{1}{2}-\frac{E_{\nu}}{m_{\mu}}\right)\,, (22)
Φν¯μ(Eν)\displaystyle\Phi_{\bar{\nu}_{\mu}}(E_{\nu}) =𝒩64Eν2mμ3(34Eνmμ),\displaystyle={\cal N}\frac{64E_{\nu}^{2}}{m_{\mu}^{3}}\left(\frac{3}{4}-\frac{E_{\nu}}{m_{\mu}}\right)\,, (23)
Φνμ(Eν)\displaystyle\Phi_{\nu_{\mu}}(E_{\nu}) =𝒩δ(Eνmπ2mμ22mπ),\displaystyle={\cal N}\delta\left(E_{\nu}-\frac{m_{\pi}^{2}-m_{\mu}^{2}}{2m_{\pi}}\right)\,, (24)

where δ\delta is the Dirac δ\delta-function, and the overall factor 𝒩rNPOT/(4πL2){\cal N}\equiv rN_{\rm POT}/(4\pi L^{2}) depends on the number of neutrinos (r)(r) that are produced for each proton on target (POT), the number of protons on target (NPOT)(N_{\rm POT}) and the distance between the source and the detector (L)(L). For the CsI[Na] detector in COHERENT experiment, r=0.0848r=0.0848, NPOT=3.198×1023N_{\rm POT}=3.198\times 10^{23} and L=19.3mL=19.3{\leavevmode\nobreak\ \rm m} Akimov et al. (2022a).

The time-integrated expected number of CEν\nuNS events in the iith bin of the number of photoelectrons (PEs) for the flavor να\nu_{\alpha} is given by Aristizabal Sierra et al. (2018b); Papoulias (2020); Atzori Corona et al. (2022); De Romeri et al. (2023)

Nναi\displaystyle N^{i}_{\nu_{\alpha}} =nNx=Cs,IηxεTναnPEinPEi+1dnPEε(nPE)\displaystyle=n_{N}\sum_{x={\rm Cs,I}}\eta_{x}\langle\varepsilon_{T}\rangle_{\nu_{\alpha}}\int_{n_{\rm PE}^{i}}^{n_{\rm PE}^{i+1}}{\rm{d}}n_{\rm PE}\varepsilon\left(n_{\mathrm{PE}}\right)
×Tnr,minTnr,maxdTnrP(nPE)dRναdTnr|x,\displaystyle\quad\times\int_{T_{\rm nr,min}}^{T_{\rm nr,max}}{\rm d}T_{\rm nr}P\left(n_{\rm PE}\right)\left.\frac{{\rm{d}}R_{\nu_{\alpha}}}{{\rm{d}}T_{\rm nr}}\right|_{x}\,, (25)

where we have taken into account the recoils of Cs and I with the fractions ηCs=51%\eta_{\rm Cs}=51\% and ηI=49%\eta_{\rm I}=49\%, respectively. The number of target nuclei in the detector is nNNAMdet/MTn_{N}\equiv N_{A}M_{\rm det}/M_{T}, where Mdet=14.6M_{\rm det}=14.6 kg is the detector active mass, MT=259.8g/molM_{T}=259.8\leavevmode\nobreak\ {\rm g/mol} is the molar mass of CsI, and NAN_{A} denotes the Avogadro number. The detector energy resolution P(nPE)P(n_{\rm PE}) and efficiency ε(nPE)\varepsilon(n_{\rm PE}) as well as the average time efficiency εTνα\langle\varepsilon_{T}\rangle_{\nu_{\alpha}} are described in Appendix A.

In Fig. 1, we compare the expected number of CEν\nuNS events in the SM as a function of nPEn_{\rm PE}, which is calculated using the nuclear response functions described in Sec. III, with the experimental data from COHERENT. The contributions from different neutrino fluxes are included.

Refer to caption
Figure 1: The comparison of the total expected event counts for CEν\nuNS events in the SM calculated using the nuclear response functions, with the experimental data (black points with error bars) collected with CsI[Na] detector by the COHERENT Collaboration Akimov et al. (2018). The numbers of events from different flavors are shown.

To constrain the neutrino NSIs, we perform the binned χ2\chi^{2} analysis using the following least-squares function Fogli et al. (2002); Aristizabal Sierra et al. (2018b),

χ2=\displaystyle\chi^{2}= i=19[NmeasiNCEνNSi(1+α)NBRN+NINi(1+β)]2(σstati)2\displaystyle\sum_{i=1}^{9}\frac{\left[N_{\rm meas}^{i}-N_{\text{CE}\nu\text{NS}}^{i}(1+\alpha)-N_{\rm BRN+NIN}^{i}(1+\beta)\right]^{2}}{(\sigma_{\rm stat}^{i})^{2}}
+(ασα)2+(βσβ)2,\displaystyle+\left(\frac{\alpha}{\sigma_{\alpha}}\right)^{2}+\left(\frac{\beta}{\sigma_{\beta}}\right)^{2}\,, (26)

where NmeasiN_{\rm meas}^{i} and NBRN+NINiN_{\rm BRN+NIN}^{i} represent the measured number of CEν\nuNS events, and the expected number of beam-related neutron (BRN) and neutrino-induced neutron (NIN) background events in the iith bin of nPEn_{\rm PE}, respectively. The associated statistical uncertainty is σstati=Nmeasi+NBRN+NINi\sigma_{\rm stat}^{i}=\sqrt{N_{\rm meas}^{i}+N_{\rm BRN+NIN}^{i}}. The expected number of CEν\nuNS events is given by

NCEνNSi=Nνei+Nνμi+Nν¯μi,\displaystyle N_{\text{CE}\nu\text{NS}}^{i}=N_{\nu_{e}}^{i}+N_{\nu_{\mu}}^{i}+N_{\bar{\nu}_{\mu}}^{i}\;, (27)

which depends on the neutrino NSIs.

The relative systematic uncertainties from the quenching factor (3.8%)(3.8\%), neutrino flux (10%)(10\%) and signal acceptance (4.1%)(4.1\%) Akimov et al. (2022a); De Romeri et al. (2023), and the response functions (5%)(5\%) are considered777It is noted that in the approach of nuclear form factor, the relative systematic uncertainty is about 5%5\% Papoulias (2020). Here, we assume that the uncertainties associated with the response functions are comparable. , which lead to the total uncertainty σα=0.125\sigma_{\alpha}=0.125. Besides, σβ=σBRN2+σNIN2\sigma_{\beta}=\sqrt{\sigma_{\rm BRN}^{2}+\sigma_{\rm NIN}^{2}} with the uncertainties of BRN and NIN backgrounds are σBRN=0.25\sigma_{\rm BRN}=0.25 and σNIN=0.35\sigma_{\rm NIN}=0.35, respectively.

The quantity χ2\chi^{2} is minimized over the systematic nuisance parameters α\alpha and β\beta, so that we can derive the 90% confidence level (C.L.) bounds on the Wilson coefficients of neutrino NSIs by requiring Δχ2χ2χmin22.71\Delta\chi^{2}\equiv\chi^{2}-\chi_{\rm min}^{2}\leq 2.71. In Eq. (8), assuming 𝒞i(6)|NSI=1\left.\mathcal{C}_{i}^{(6)}\right|_{\rm NSI}=1 and summing over the fluxes of νμ\nu_{\mu} and ν¯μ\bar{\nu}_{\mu}, we obtain the one-parameter-a-time lower bounds on ΛNSI\Lambda_{\rm NSI} for specific flavors, which are presented in Table 1.

ΛNSI/GeV\Lambda_{\rm NSI}/{\rm GeV} νe\nu_{e} νμ\nu_{\mu}
𝒞^1,u(6)\hat{\mathcal{C}}_{1,u}^{(6)} 390 395
𝒞^1,d(6)\hat{\mathcal{C}}_{1,d}^{(6)} 407 414
𝒞^2,u(6)\hat{\mathcal{C}}_{2,u}^{(6)} 44.7 68.4
𝒞^2,d(6)\hat{\mathcal{C}}_{2,d}^{(6)} 26.4 40.9
Table 1: Lower bounds on ΛNSI\Lambda_{\rm NSI} in units of GeV for the Wilson coefficients 𝒞^i(6)|NSI=𝒞^i(6)𝒞^i(6)|SM\left.\hat{\mathcal{C}}_{i}^{(6)}\right|_{\rm NSI}=\hat{\mathcal{C}}_{i}^{(6)}-\left.\hat{\mathcal{C}}_{i}^{(6)}\right|_{\rm SM} from COHERENT Akimov et al. (2022a).

IV.2 Constraints from PandaX-4T

The recent measurements of solar 8B neutrinos in the CEν\nuNS process Ma et al. (2023); Bo et al. (2024); Aprile et al. (2024) can also impose constraints on the neutrino NSIs. In this work, we consider the results of the PandaX-4T experiment using the liquid xenon Bo et al. (2024).

The differential event rate per target for neutrino flavor να=νe,νμ,ντ\nu_{\alpha}=\nu_{e},\nu_{\mu},\nu_{\tau} is expressed as

dRναdTnr=Eν,minEν,max𝑑EνΦνα(Eν)dσdTnr,\frac{{{\rm d}}R_{\nu_{\alpha}}}{{{\rm d}}T_{\rm nr}}=\int_{E_{\nu,{\rm min}}}^{E_{\nu,{\rm max}}}dE_{\nu}\Phi_{\nu_{\alpha}}(E_{\nu})\frac{{\rm d}\sigma}{{\rm d}T_{\rm nr}}\,, (28)

where the minimum neutrino energy Eν,minMTnr/2E_{\nu,\rm min}\simeq\sqrt{MT_{\rm nr}/2} with MM the mass of 131Xe, and the maximum energy Eν,maxE_{\nu,\rm max} is about 16 MeV\rm MeV Bahcall et al. (1996).

The solar 8B neutrino νe\nu_{e} is produced in the Sun, and then propagates to the Earth. The total flux of neutrino να\nu_{\alpha} detected at the Earth is defined as

Φνα(Eν)=MdetPναϕ(8B),\displaystyle\Phi_{\nu_{\alpha}}(E_{\nu})=\dfrac{\mathcal{E}}{M_{\rm det}}\langle P_{\nu_{\alpha}}\rangle\phi(^{8}{\rm B})\;, (29)

where \mathcal{E} and MdetM_{\rm det} denote the exposure and detector active mass, respectively, and ϕ(8B)=5.46(1±0.12)×106cm2s1\phi(^{8}{\rm B})=5.46(1\pm 0.12)\times 10^{6}\leavevmode\nobreak\ {\rm cm^{-2}s^{-1}} is the predicted solar 8B neutrino flux Vinyoles et al. (2017)888The predictions by the other groups based on the standard solar model can be found in Refs. Bahcall et al. (1996, 2006). . Pνα\langle P_{\nu_{\alpha}}\rangle is the probability of solar neutrino νe\nu_{e} to manifest as να\nu_{\alpha} at the Earth averaged over the exposure.

Due to the neutrino oscillation, the flavor composition of solar neutrinos detected at the Earth differs from that produced in the Sun. In the analysis, we use the package PEANUTS Gonzalo and Lucente (2024) to compute Pνα\langle P_{\nu_{\alpha}}\rangle with matter effects in the neutrino propagation being included, which are shown in Fig. 2 for different flavors of neutrinos.

Refer to caption
Figure 2: The probabilities of various neutrino fluxes at the Earth, averaged over the exposure period, are detailed.

The expected number of CEν\nuNS events for the flavor να\nu_{\alpha} is given by Li et al. (2024),

Nνα=nNTnr,minTnr,maxdTnrε(Tnr)dRαdTnr,\displaystyle N_{\nu_{\alpha}}=n_{N}\int^{T_{\rm nr,max}}_{T_{\rm nr,min}}{\rm{d}}T_{\rm nr}\leavevmode\nobreak\ \varepsilon(T_{\rm nr})\frac{{\rm{d}}R_{\alpha}}{{\rm{d}}T_{\rm nr}}\,, (30)

where ε(Tnr)\varepsilon(T_{\rm nr}) is the detection efficiency, which depends on the nuclear recoil function. The number of target nuclei (131Xe)(^{131}\rm Xe) in the detector of the PandaX-4T experiment is nNNAMdet/MTn_{N}\equiv N_{A}M_{\rm det}/M_{T} with the molar mass MT=131.29g/molM_{T}=131.29\leavevmode\nobreak\ {\rm g/mol}.

Two datasets are collected by PandaX-4T, which differ in the energy threshold, as displayed in Table 2. Note that the numbers of signal events for the pair data (3.5)(3.5) and US2 data (75)(75) are obtained from the combined likelihood fit Bo et al. (2024). Given the smaller number of signal events in paired data, we only consider the US2 data. The exposure in Eq. (29) is thus given by =1.04tonneyear\mathcal{E}=1.04\leavevmode\nobreak\ {\rm tonne}\cdot{\rm year}.

PandaX-4T data paired US2
energy threshold 1.11.1 keV 0.330.33 keV
total exposure 1.251.25 tonne\cdotyear 1.041.04 tonne\cdotyear
event number 3.5 75
Table 2: The experimental data of detecting solar 8B neutrinos in the CEν\nuNS process by PandaX-4T Bo et al. (2024). The energy thresholds, total exposures, and numbers of signal events for the paired and US2 data are given.

In the analysis by the PandaX-4T Collaboration Bo et al. (2024), the signals are interpreted in terms of the measured solar 8B neutrino flux assuming the SM cross section of CEν\nuNS, which is (8.4±3.1)×106cm2(8.4\pm 3.1)\times 10^{6}\mathrm{\leavevmode\nobreak\ cm}^{-2} with the relative statistical uncertainty being 37%37\%. Therefore, for our purpose, instead of taking the number of signal events post the combined likelihood fit, we calculate the number of signal events for the US2 data using the signal efficiency from Figure 1 of Ref. Bo et al. (2024). We obtain the measured number of CEν\nuNS events Nmeas=69.1N_{\rm meas}=69.1 with the statistic uncertainty σstat=0.37×Nmeas\sigma_{\rm stat}=0.37\times N_{\rm meas}.

To constrain the neutrino NSIs, we perform the single-bin χ2\chi^{2} analysis Fogli et al. (2002),

χ2=[NmeasNCEνNS(1+α)]2σstat2+(ασα)2.\chi^{2}=\frac{\left[N_{\rm meas}-N_{\text{CE}\nu\text{NS}}(1+\alpha)\right]^{2}}{\sigma_{\rm stat}^{2}}+\left(\frac{\alpha}{\sigma_{\alpha}}\right)^{2}\,. (31)

Here, the expected number of CEν\nuNS events is given by

NCEνNS=Nνe+Nνμ+Nντ,\displaystyle N_{\text{CE}\nu\text{NS}}=N_{\nu_{e}}+N_{\nu_{\mu}}+N_{\nu_{\tau}}\;, (32)

which depends on the neutrino NSIs, and the SM prediction NCEνNS|SM=44.9\left.N_{\text{CE}\nu\text{NS}}\right|_{\rm SM}=44.9 calculated using the solar 8B neutrino flux ϕ(8B)=5.46(1±0.12)×106cm2s1\phi(^{8}{\rm B})=5.46(1\pm 0.12)\times 10^{6}\leavevmode\nobreak\ {\rm cm^{-2}s^{-1}} Vinyoles et al. (2017). The relative systematic uncertainties from the selection efficiency (12%)(12\%), signal modeling (17%)(17\%) and solar 8B neutrino flux (12%)(12\%) Bo et al. (2024); Vinyoles et al. (2017), and the response functions (5%)(5\%) are considered, which lead to the total uncertainty σα=0.245\sigma_{\alpha}=0.245.

Again, χ2\chi^{2} is minimized over the nuisance parameter α\alpha to derive the 90% C.L. limits on the Wilson coefficients by requiring Δχ2χ2χmin22.71\Delta\chi^{2}\equiv\chi^{2}-\chi_{\rm min}^{2}\leq 2.71. In Eq. (8), assuming 𝒞i(6)|NSI=1\left.\mathcal{C}_{i}^{(6)}\right|_{\rm NSI}=1, we obtain the one-parameter-a-time lower bounds on ΛNSI\Lambda_{\rm NSI} for specific flavors, which are presented in Table 3.

ΛNSI/GeV\Lambda_{\rm NSI}/{\rm GeV} νe\nu_{e} νμ\nu_{\mu} ντ\nu_{\tau}
𝒞^1,u(6)\hat{\mathcal{C}}_{1,u}^{(6)} 287.46 289.61 286.70
𝒞^1,d(6)\hat{\mathcal{C}}_{1,d}^{(6)} 304.61 306.88 303.80
𝒞^2,u(6)\hat{\mathcal{C}}_{2,u}^{(6)} 14.70 14.81 14.60
𝒞^2,d(6)\hat{\mathcal{C}}_{2,d}^{(6)} 23.32 23.48 23.15
Table 3: Lower bounds on ΛNSI\Lambda_{\rm NSI} in units of GeV for the Wilson coefficients 𝒞^i(6)|NSI=𝒞^i(6)𝒞^i(6)|SM\left.\hat{\mathcal{C}}_{i}^{(6)}\right|_{\rm NSI}=\hat{\mathcal{C}}_{i}^{(6)}-\left.\hat{\mathcal{C}}_{i}^{(6)}\right|_{\rm SM} from PandaX-4T Bo et al. (2024).

Owing to the neutrino oscillation, a significant portion of the solar neutrino fluxes reaching the Earth are composed of ντ\nu_{\tau} as depicted in Fig. 2. Therefore, the CEν\nuNS measurement of PandaX-4T can give unique constraints on the neutrino NSIs for the τ\tau flavor as shown in the last column of Table 3. For the ee and μ\mu flavors, the measurement of COHERENT CsI provides more stringent constraints on the Wilson coefficients of NSIs than the constraints provided by PandaX-4T.

For comparison, we also obtain the two-dimensional constraint on the neutrino NSI parameters ϵeeuV\epsilon_{ee}^{uV} and ϵeedV\epsilon_{ee}^{dV}, which are defined as Abdullah et al. (2022)

ϵeeqV=12GF𝒞1,q(6)|NSI.\displaystyle\epsilon_{ee}^{qV}=\dfrac{-1}{\sqrt{2}G_{F}}\left.\mathcal{C}_{1,q}^{(6)}\right|_{\rm NSI}\;. (33)

By requiring Δχ24.61\Delta\chi^{2}\leq 4.61, we obtain the 90% C.L. allowed regions in Fig. 3. Note that our fitted result using the COHERENT CsI measurement agrees with Ref. De Romeri et al. (2023). It is shown that the constraints on ϵeeuV\epsilon_{ee}^{uV} and ϵeedV\epsilon_{ee}^{dV} from the CEν\nuNS measurement by PandaX-4T are weaker than those from COHERENT.

Refer to caption
Figure 3: The 90% C.L. allowed regions of ϵeeuV\epsilon_{ee}^{uV} and ϵeedV\epsilon_{ee}^{dV} by the measurements in the COHERENT CsI Akimov et al. (2022a) and PandaX-4T Bo et al. (2024) experiments, as depicted in blue and red colors, respectively.
Refer to caption
Figure 4: The distribution of Δχ2\Delta\chi^{2} as a function of the CEν\nuNS counts. The SM prediction of the CEν\nuNS counts is 44.9, located at the endpoint of the black dotted line. The central value of measured CEν\nuNS counts in the PandaX-4T experiment (US2 data) is 69.169.1, located at the minimum of Δχ2\Delta\chi^{2} distribution. The gray line corresponds to Δχ2=4.61\Delta\chi^{2}=4.61.
Refer to caption
Figure 5: The 90% C.L. allowed regions of ϵeeuV\epsilon_{ee}^{uV} and ϵeedV\epsilon_{ee}^{dV} by the COHERENT CsI measurement (blue) and anticipated measurements by PandaX-4T assuming the experimental central value Nmeas=NCEνNS|SM=44.9N_{\rm meas}=\left.N_{\text{CE}\nu\text{NS}}\right|_{\rm SM}=44.9 and exposure =6tonneyear\mathcal{E}=6\leavevmode\nobreak\ {\rm tonne}\cdot{\rm year}. The lighter and darker red regions are obtained with the systematic uncertainty σα=0.245\sigma_{\alpha}=0.245 and 0, respectively.

We find that the central value of experimentally measured CEν\nuNS counts has a significant impact on the constraints on the NSI parameters999It is noted that the central value measured by XENONnT is slightly smaller than the SM prediction, whereas the associated statistical uncertainty is larger compared to PandaX-4T Aprile et al. (2024).. To understand it, we present the Δχ2\Delta\chi^{2} distribution as a function of the CEν\nuNS counts in Fig. 4, using the US2 data from PandaX-4T. The upper limit of the CEν\nuNS counts with Δχ24.61\Delta\chi^{2}\leq 4.61 determines the boundaries of the blue bands in Fig. 5. It is verified that with the central value unchanged, the sensitivity of PandaX-4T shows a mild improvement for 5\sim 5 times larger exposure.

We further assess the sensitivities of future solar 8B neutrino measurements via CEν\nuNS in DM detectors, leveraging the current results from PandaX-4T as a reference. In Fig. 5, we show the projected constraints on the NSI parameters under the assumption of Nmeas=NCEνNS|SM=44.9N_{\rm meas}=\left.N_{\text{CE}\nu\text{NS}}\right|_{\rm SM}=44.9 and =6tonneyear\mathcal{E}=6\leavevmode\nobreak\ {\rm tonne}\cdot{\rm year}. We consider two cases with the total systematic uncertainty σα=0.245\sigma_{\alpha}=0.245 being included or not. The resulting constraints are represented in lighter red and darker red bands with solid and dashed boundaries, respectively. One can observe that the anticipated constraint from PandaX-4T with the systematic uncertainties is weaker than that from COHERENT. If in the future the systematic uncertainties are well controlled, with the assumption that they are negligible, we could achieve a sensitivity comparable to that of the COHERENT CsI measurement.

V Conclusion

In this work, we have studied the constraints on the neutrino non-standard interactions (NSIs) using the latest CEν\nuNS measurements in the COHERENT and PandaX-4T experiments. The cross section of CEν\nuNS is calculated within an end-to-end effective field theory framework. In this approach, the dimension-6 operators in the low-energy EFT are matched to hadronic operators in the heavy baryon chiral perturbation theory, while the nuclear response of target nuclei is described using multipole expansions.

We have performed the χ2\chi^{2} analyses of the CEν\nuNS events observed in the CsI[Na] detector of COHERENT and PandaX-4T, and have derived the one-parameter-a-time lower bounds on the NSI scale for specific neutrino flavors. We have found that the constraints for the ee and μ\mu flavors from COHERENT are more stringent than those from PandaX-4T, while the latter provides unique sensitivities to the neutrino NSIs for the τ\tau flavor due to the oscillation of solar 8B neutrinos propagating from the Sun to the Earth (considering the matter effects in the neutrino propagation by using package PEANUTS). Besides, we have compared the two-dimensional constraints on the NSI parameters for the ee flavor from COHERENT and PandaX-4T, and have obtained that the sensitivity of PandaX-4T is limited by the central value of measured CEν\nuNS counts.

Moreover, we have assessed the measurements of the solar 8B neutrinos via CEν\nuNS in dark matter detectors, leveraging the current results from PandaX-4T as a reference. Assuming that the central value of measured CEν\nuNS counts aligns with the SM prediction, the sensitivity is significantly improved for the exposure of 6tonneyear6\leavevmode\nobreak\ {\rm tonne}\cdot{\rm year}, and is comparable to that imposed by the COHERENT CsI measurement if the systematic uncertainties are further disregarded.

Note added: After this paper was finished, another paper Aristizabal Sierra et al. (2024) appeared, which has some overlap of our work. However, we use an end-to-end EFT framework to investigate the neutrino NSIs sensitivities, and we emphasize that it is important to include systematic uncertainties in the analysis.

Acknowledgements.
We would like to express our gratitude to Yu-Feng Li for the valuable help with the χ2\chi^{2} fitting of the COHERENT data. GL also thanks Xun-Jie Xu for useful correspondence regarding Ref. Lindner et al. (2017). FT thanks Ningqiang Song for the discussion on the solar neutrino and Bing-Long Zhang for the discussion on numerical calculations. This work is supported by the National Science Foundation of China under Grants No. 12347105, No. 12375099 and No. 12047503, and the National Key Research and Development Program of China Grant No. 2020YFC2201501, No. 2021YFA0718304. GL is also supported by the Guangdong Basic and Applied Basic Research Foundation (2024A1515012668), and SYSU startup funding.

Appendix A Detector resolution and efficiency in COHERENT CsI measurement

In Eq. (IV.1), the number of PEs is Papoulias (2020); Akimov et al. (2022a)

nPE=13.35TeekeV,\displaystyle n_{\mathrm{PE}}=13.35\leavevmode\nobreak\ \dfrac{T_{ee}}{\rm keV}\;, (34)

where the true electron-equivalent recoil energy TeeT_{ee} is related to the true nuclear recoil energy as

Tee=fQ(Tnr)Tnr.\displaystyle T_{ee}=f_{Q}(T_{\rm nr})T_{\rm nr}\;. (35)

The quenching factor fQ(Tnr)f_{Q}(T_{\rm nr}) can be parameterized as Akimov et al. (2022b)

fQ(Tnr)=k0+k1Tnr+k2Tnr2+k3Tnr3,\displaystyle f_{Q}(T_{\rm nr})=k_{0}+k_{1}T_{\rm nr}+k_{2}T_{\rm nr}^{2}+k_{3}T_{\rm nr}^{3}\;, (36)

where the parameters k0=0.05546k_{0}=0.05546, k1=4.307k_{1}=4.307, k2=111.7k_{2}=-111.7 and k3=840.4k_{3}=840.4.

The detector energy resolution P(nPE)P(n_{\rm PE}) is modeled with the gamma function,

P(nPE)=(a(1+b))1+bΓ(1+b)nPEbea(1+b)nPE,\displaystyle P(n_{\rm PE})=\frac{(a(1+b))^{1+b}}{\Gamma(1+b)}n_{\rm PE}^{b}e^{-a(1+b)n_{\rm PE}}\;, (37)

where a=0.0749keV/Eeea=0.0749{\leavevmode\nobreak\ \rm keV}/E_{ee} and b=9.56Eee/keVb=9.56E_{ee}/{\rm keV}.

The reconstructed energy and time are uncorrelated, thus allowing us to deal with the energy and time efficiency independently Akimov et al. (2022a),

εE(nPE)\displaystyle\varepsilon_{E}(n_{\rm PE}) =a11+eb1(nPEc1)+d1,\displaystyle=\frac{a_{1}}{1+e^{-b_{1}(n_{\rm PE}-c_{1})}}+d_{1}\;, (38)
εT(trec)\displaystyle\varepsilon_{T}\left(t_{\rm rec}\right) ={1,trec<a2,eb2(treca2),treca2,\displaystyle=\begin{cases}1\;,&t_{\rm rec}<a_{2}\;,\\ e^{-b_{2}\left(t_{\rm rec}-a_{2}\right)}\;,&t_{\rm rec}\geq a_{2}\;,\end{cases} (39)

where the parameters are a1=1.32a_{1}=1.32, b1=0.285b_{1}=0.285, c1=10.9c_{1}=10.9, d1=0.333d_{1}=-0.333, and a2=0.52μa_{2}=0.52\leavevmode\nobreak\ \mus, b2=0.0494/μb_{2}=0.0494/\muAkimov et al. (2022a).

The efficiency ε(nPE)\varepsilon(n_{\rm PE}) has been implemented in the integration over nPEn_{\rm PE} in Eq. (IV.1). On the other hand, we consider the average time efficiency

εTναjtrecitrecj+1dtrecεT(trec)Nναj(trec)jtrecitrecj+1dtrecNναj(trec),\displaystyle\langle\varepsilon_{T}\rangle_{\nu_{\alpha}}\equiv\dfrac{\sum_{j}\int_{t_{\rm rec}^{i}}^{t_{\rm rec}^{j+1}}{\rm d}t_{\rm rec}\leavevmode\nobreak\ \varepsilon_{T}(t_{\rm rec})N^{j}_{\nu_{\alpha}}(t_{\rm rec})}{\sum_{j}\int_{t_{\rm rec}^{i}}^{t_{\rm rec}^{j+1}}{\rm d}t_{\rm rec}N^{j}_{\nu_{\alpha}}(t_{\rm rec})}\;, (40)

where Nναj(trec)N^{j}_{\nu_{\alpha}}(t_{\rm rec}) represents the nPEn_{\rm PE}-integrated expected number of the SM CEν\nuNS events in the reconstructed time trect_{\rm rec} for each flavor of neutrino flux depicted in the right panel of Figure 1 in Ref. Akimov et al. (2022a). Due to variations in neutrino arrival times, the time efficiencies for different flavors are distinct, ϵTνμ=0.994\langle\epsilon_{T}\rangle_{\nu_{\mu}}=0.994, ϵTν¯μ=0.918\langle\epsilon_{T}\rangle_{\bar{\nu}_{\mu}}=0.918 and ϵTνe=0.92\langle\epsilon_{T}\rangle_{\nu_{e}}=0.92.

References