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

The Sun: light dark matter and sterile neutrinos

Ilídio Lopes,21{}^{1},^{2} ID [ 1Centro de Astrofísica e Gravitação - CENTRA,
Departamento de Física, Instituto Superior Técnico - IST, Universidade de Lisboa - UL, Av. Rovisco Pais 1, 1049-001 Lisboa, Portugal
2 Institut d’Astrophysique de Paris, UMR 7095 CNRS, Université Pierre et Marie Curie, 98 bis Boulevard Arago, Paris F-75014, France
Abstract

Next-generation experiments allow for the possibility to testing the neutrino flavor oscillation model to very high levels of accuracy. Here, we explore the possibility that the dark matter in the current universe is made of two particles, a sterile neutrino and a very light dark matter particle. By using a 3+1 neutrino flavor oscillation model, we study how such a type of dark matter imprints the solar neutrino fluxes, spectra, and survival probabilities of electron neutrinos. The current solar neutrino measurements allow us to define an upper limit for the ratio of the mass of a light dark matter particle mϕm_{\phi} and the Fermi constant GϕG_{\phi}, such that Gϕ/mϕG_{\phi}/m_{\phi} must be smaller than 1030GFeV110^{30}\;G_{\rm F}eV^{-1} to be in agreement with current solar neutrino data from the Borexino, Sudbury Neutrino Observatory, and Super-Kamiokande detectors. Moreover, for models with a very small Fermi constant, the amplitude of the time variability must be lower than 3%3\% to be consistent with current solar neutrino data. We also found that solar neutrino detectors like Darwin, able to measure neutrino fluxes in the low energy-range with high accuracy, will provide additional constraints to this class of models that complement the ones obtained from the current solar neutrino detectors.

Subject headings:
The Sun – Solar neutrino problem – Solar neutrinos – Neutrino oscillations – Neutrino telescopes – Neutrino astronomy

1. Introduction

The origin of dark matter has been a fundamental problem in physics for almost six decades, during which most of the proposed solutions assumed that a single massive particle that interacts weakly with baryons makes all the dark matter observed in the universe (e.g., Wang et al., 2016). Recently, research has emerged where more sophisticated solutions have been proposed to solve the dark matter problem. One of these is the possibility of the dark matter being a composite of two light particles: a light dark matter (LDM) particle ϕ\phi and a sterile neutrino νs\nu_{s}.

The existence of such an LDM field can by identified with a dilation field of an extradimensional extension of the Standard Model or/and a CP-violating pseudo-Goldstone boson of a spontaneously broken global symmetry. For some of these models, ϕ\phi couples to the Standard Model fields, and as such it induces periodic time variation in particle masses and couplings. In such theories the gauge invariance suggests that the ϕ\phi should possess an identical coupling constant to charged leptons, in which case scalar interactions with the electrons provide a good opportunity for detection through atomic clocks (e.g., Arvanitaki et al., 2016), accelerometers (e.g., Arvanitaki et al., 2018), and gravitational wave detectors (e.g., Lopes and Silk, 2014; Graham et al., 2016).

Similarly, this ϕ\phi field can couple to neutrinos. Once again, these types of interactions generically result in time-varying corrections to the neutrino masses, neutrino mass differences, and mixing angles, which can be searched for in the neutrino flux signals on present and future experimental neutrino detectors (Aharmim et al., 2013; Abe et al., 2016; Borexino Collaboration et al., 2018; Aalbers et al., 2020). If ϕ\phi couples weakly to the neutrinos, over a large range of masses, it can significantly modify the neutrino oscillations probabilities leading to a distorted survival electron neutrino probability function (Berlin, 2016; Krnjaic et al., 2018).

The motivation for such a model comes from the possibility of this composite particle physics model resolving two observational problems:

  1. 1.

    The classical cold dark matter model leads to several inconsistencies with the cosmological observational data, such as the missing satellite problem and the cusp problem (e.g., Primack, 2009). Light dark matter resolves such problems if dark matter is totally or partially made of light scalar particles with a mass of the order of 1022eV10^{-22}{\rm eV} (Hu et al., 2000; Peebles, 2000). In hierarchical models of structure formation, such type of dark matter is able to explain the flatness observed on the profiles of the distribution of gas and stars in halos and filaments (Mocz et al., 2019).

  2. 2.

    Although the standard three neutrino flavor model produces a reasonable good global fit to all the neutrino data (Esteban et al., 2019), there are now many hints that point out to the possibility of the existence of a fourth neutrino. This one does not have any other interaction than gravity and for that reason is known as a sterile neutrino (Diaz et al., 2019). It was found that a flavor oscillation model made of the 3 active neutrinos plus a sterile neutrino could explain some of the observed anomalies found on the Short baseline neutrino oscillation experiments (Giunti et al., 2012, 2013), Liquid Scintillator Neutrino Detector (Aguilar et al., 2001), and MiniBooNE Short-Baseline Neutrino Experiment (MiniBooNE Collaboration et al., 2018), as well as the anomalies related with GALLEX and SAGE solar-neutrino detectors – the so-called Gallium anomalies (Kostensalo et al., 2019). For instance, Kostensalo et al. (2019) found that the data favour a 3+13+1 neutrino flavor model with m4= 1.1eVm_{4}=\;1.1eV and mixing matrix element Ue4=0.11U_{e4}=0.11.

One possibility to resolve both problems (neutrino anomalies and structure formation) is to consider that dark matter is made of a light scalar field that couples to a sterile neutrino (e.g., Farzan, 2019). The interactions of ϕ\phi and νs\nu_{s} could impede the oscillations in the universe and thereby improve the agreement between the structure formation and cosmological observations (e.g., Dasgupta and Kopp, 2014; Hannestad et al., 2014).

If such ϕ\phi and νs\nu_{s} particles exist today, they were produced abundantly in the early universe. For instance, sterile neutrinos can be produced via mixing with active neutrinos (Dodelson and Widrow, 1994), in some scenarios such neutrino production is being enhanced by the oscillations between active and sterile neutrinos (Bezrukov et al., 2019, 2020; de Gouvêa et al., 2020) or by the lepton asymmetry (Shi and Fuller, 1999). The production of light dark matter can take many forms, such as vector bosons by parametric resonance production (Dror et al., 2019). For instance, some models predict a sterile neutrino abundance of Ωs2h2=0.12(sin2(2θs)/3.5×109)(mνs/7keV)\Omega_{s}^{2}h^{2}=0.12\left({\sin^{2}{(2\theta_{s})}}/{3.5\times 10^{-9}}\right)\left({m_{\nu_{s}}}/{7keV}\right) where mνsm_{\nu_{s}} and θs\theta_{s} is the sterile neutrino mass and mixing angle (Kusenko, 2009). For the light dark matter field some authors obtained Ωϕ2=0.1(ao/1017GeV)2(ma/1022eV)1/2\Omega_{\phi}^{2}=0.1\left({a_{o}}/{10^{17}\;{\rm GeV}}\right)^{2}\left({m_{a}}/{10^{-22}\;{\rm eV}}\right)^{1/2}, where aoa_{o} is a parameter that relates to the initial misalignment angle of the axion, and mam_{a} is the axion mass (Hui et al., 2017; Niemeyer, 2019). Conveniently, we will assume that in the present-day universe the total dark matter abundance is given by

ΩDMh2=Ωϕ2h2+Ωνs2h2\displaystyle\Omega_{\rm DM}h^{2}=\Omega_{\phi}^{2}h^{2}+\Omega_{\nu_{s}}^{2}h^{2} (1)

where the Ωϕ2h2\Omega_{\phi}^{2}h^{2} and Ωs2h2\Omega_{s}^{2}h^{2} are the total ϕ\phi and νs\nu_{s} densities in the present universe, respectively. For future reference, we assume that the present-day total dark matter abundance ΩDMh2=0.12\Omega_{\rm DM}h^{2}=0.12 (Planck Collaboration et al., 2018), and the dark matter density in the solar neighborhood is ρDM=0.39GeVcm3\rho^{\odot}_{\rm DM}=0.39\;GeV\;cm^{-3} (Catena and Ullio, 2010).

In this paper, we study the impact that this light dark matter field has in the 3+1 neutrino flavor model. Specifically, we discuss how the light dark matter field modifies the neutrino flavor oscillations, and by using the current sets of solar neutrino data, we also put constraints in the parameters of such models and make predictions for the future neutrino experiments.

The article is organized as follows. In section 2, we discuss how the light dark matter drives the 3+1 neutrino flavor oscillations. In section 3, we present the neutrino flavor oscillation model in the presence of a cosmic light dark matter field. In section 4, we compute the survival electron neutrino probabilities for the electron neutrinos produced in the proton-proton (PP) chain and carbon-nitrogen-oxygen (CNO) cycle solar nuclear reactions. In section 5, we discuss the results in relation to current experiments and future ones. Finally, in section 6, we present the conclusion and a summary of our results.

If not stated otherwise, we work in natural units in which c==1c=\hbar=1. In these units all quantities are measured in GeV, and we make use of the conversion rules 1m=5.068×1015GeV11m=5.068\times 10^{15}GeV^{-1}, 1kg=5.610×1026GeV1kg=5.610\times 10^{26}GeV and 1sec=1.519×1024GeV11sec=1.519\times 10^{24}GeV^{-1}.

2. Light Dark Matter and Sterile Neutrinos in the Universe

We assume that in the present universe, the dark matter is composed of two fundamental particles: a light scalar boson ϕ\phi and sterile neutrinos νs\nu_{s}, where mϕm_{\phi} and mνsm_{\nu_{s}} are their respective masses (Hannestad et al., 2014). The LDM field ϕ\phi couples with the active neutrinos and the sterile neutrino by a Yukawa interaction gϕϕνsνsg_{\phi}\phi\nu_{s}\nu_{s} where gϕg_{\phi} is a dimensionless coupling (Farzan, 2019). To illustrate this effect, consider an LDM scalar ϕ\phi with a Yukawa coupling to active neutrinos. Then the relevant part of the Lagrangian reads

(mν+gϕ)νν+H.c.,\displaystyle{\cal L}\supset-(m_{\nu}+g\phi)\nu\nu+H.c., (2)

where for convenience of representation the flavor indices have been suppressed. We also assume that the dimensionless coupling is very small (g1g\ll 1). From the Euler-Lagrange equations of ν\nu and ϕ\phi, it is possible to show that the effect of ϕ\phi on the propagation of the neutrino is equivalent to changing the neutrino mass from mνm_{\nu} to mν+δmνm_{\nu}+\delta m_{\nu}. As we will see later, this δmν\delta m_{\nu} perturbation will induce time variations in the mass-squared differences and mixing angles of all neutrino flavors through ϕ\phi. In principle the Yukawa couplings can have any structure in the neutrino flavor space. In this work, we will focus on two convenient scenarios of great interest to neutrino detectors: mass-square differences and mixing angles (e.g., Ding and Feruglio , 2020). Moreover, we will also assume that δmν=gϕϕ\delta m_{\nu}=g_{\phi}\phi (e.g., Smirnov and Xu, 2019).

2.1. Dark Matter Time-dependent Variation

The hypothesis that dark matter in the local universe is made of very light particles leads to the following description: the LDM field ϕ\phi in the dark matter halo of the Milky Way is represented by a group of plane waves with frequency ωϕ\omega_{\phi}, such that ωϕ=mϕ(1+vϕ2/2)\omega_{\phi}=m_{\phi}(1+v_{\phi}^{2}/2) where vϕv_{\phi} is the virial velocity of the particles in the dark halo. A population of such light particles will smooth inhomogeneities in the dark matter distribution on scales smaller than the de Broglie wavelength λdB\lambda_{\rm dB} of these LDM particles. For any particle, we compute λdB\lambda_{\rm dB} using the relation λdB=1.24×1022(1023eV/mϕ)(103/vϕ)cm\lambda_{\rm dB}=1.24\times 10^{22}\left(10^{-23}{\rm eV}/m_{\phi}\right)\left(10^{-3}/v_{\phi}\right)\;{\rm cm}.

We notice that the kinetic term on ωϕ\omega_{\phi} is neglected once the virial velocity vϕ103v_{\phi}\sim 10^{-3} is very small (e.g., Blas et al., 2017). Therefore, we dropped the corrections related with vϕv_{\phi} for the equation 2. Accordingly, the general form of this LDM field reads

ϕ(r,t)=ϕocos(mϕt+ϵo)ϕocos(mϕt),\displaystyle\phi(\vec{r},t)=\phi_{o}\cos{(m_{\phi}t+\epsilon_{o})}\approx\phi_{o}\cos{(m_{\phi}t)}, (3)

where ϕo\phi_{o} and ϵo=vϕr\epsilon_{o}=\vec{v}_{\phi}\cdot\vec{r} are the amplitude and phase of the wave ϕ(t)\phi(t), respectively. In this work we consider ϵo0\epsilon_{o}\approx 0. Moreover, the energy momentum of a free massive oscillating field has a density given by ρϕ=ϕomϕ/2\rho_{\phi}=\phi_{o}m_{\phi}/2 and a pressure given by pϕ=ρϕcos(2mϕt)p_{\phi}=-\rho_{\phi}\cos{(2m_{\phi}t)}. Although formally ρϕ\rho_{\phi} has an oscillating part proportional to ϕ(t)\phi(t), because this component is very small we neglected its contribution in this analysis (Khmelnitsky and Rubakov, 2014).

The quantity ϕo\phi_{o} is a slowly varying function of the position. Conveniently, the amplitude of ϕ(t)\phi(t) can be written as ϕo=2ρϕ(r)/mϕ\phi_{o}=\sqrt{2\rho_{\phi}(r)}/m_{\phi} where ρϕ(r)=ρDM(Ωϕ/ΩDM)\rho_{\phi}(r)=\rho^{\odot}_{\rm DM}\left(\Omega_{\phi}/\Omega_{\rm DM}\right) is the fraction of dark matter density in ϕ\phi particles at the space-time coordinate r\vec{r}. Accordingly, the Sun immersed in this light dark matter halo will experience a periodic perturbation due to the action of the ϕ(r,t)\phi(\vec{r},t), which by the presence of a Yukawa coupling gϕg_{\phi} will exert a temporal variation on the propagation of all neutrinos. We estimate the dark matter density number nϕn_{\phi} in the solar neighborhood as follows: if we consider that the main contribution arises from a single dark matter particle with mass mϕm_{\phi}, then the relevant density in our case will take the value nϕ=ρϕ/mϕn_{\phi}=\rho_{\phi}/m_{\phi}. If we assume that all dark matter is made of ϕ\phi bosons, we have ρϕ=ρDM=0.39GeVcm3\rho_{\phi}=\rho^{\odot}_{\rm DM}=0.39\;GeV\;cm^{-3} (Catena and Ullio, 2010) and mϕ=1022eVm_{\phi}=10^{-22}\;{\rm eV} then nϕ=3.9×1030cm3n_{\phi}=3.9\times 10^{30}\;cm^{-3} (particles per centimeter cubed). This value is only 2 orders of magnitude smaller than the density of electrons in the Sun’s core, ne6 1031cm3n_{e}\sim 6\;10^{31}\;cm^{-3} (Lopes and Turck-Chièze, 2013). Since these particles are very light, we assume that there is no accretion of these particles in the Sun’s core during its evolution in the main sequence until the present age.

2.2. Neutrino Time-dependent Dark-matter-induced Oscillations

In the presence of the LDM field ϕ\phi, the neutrino mass mνm_{\nu}, according to equation 2 (Ding and Feruglio , 2020), will receive a contribution δmν=gϕ\delta m_{\nu}=g\phi, such that from equation 3, we obtain

δmνmν=ϵϕcos(mϕt),\displaystyle\frac{\delta m_{\nu}}{m_{\nu}}=\epsilon_{\phi}\cos{(m_{\phi}t)}, (4)

where ϵϕ\epsilon_{\phi} is the amplitude

ϵϕ=gϕ2ρϕmϕmν=gϕ2ρDMmϕmν(ΩϕΩDM)1/2.\displaystyle\epsilon_{\phi}=\frac{g_{\phi}\sqrt{2\rho_{\phi}}}{m_{\phi}m_{\nu}}=\frac{g_{\phi}\sqrt{2\rho^{\odot}_{\rm DM}}}{m_{\phi}m_{\nu}}\;\left(\frac{\Omega_{\phi}}{\Omega_{\rm DM}}\right)^{1/2}. (5)

If not stated otherwise, we will assume that all dark matter in the present universe is made of only LDM particles such that Ωϕ=ΩDM\Omega_{\phi}=\Omega_{\rm DM}. We observe that ϵϕ\epsilon_{\phi} is a relevant factor even if ϕ\phi is a small fraction of the dark matter halo. In particular, ϕ\phi will affect the oscillation parameters of all neutrino flavors, including the sterile sterline neutrinos. If we only take into account the first order perturbation, thus, the neutrino mass-squared difference can be written as

Δmij2(t)=mi2mj2Δmij,o2[1+2ϵϕcos(mϕt)]\displaystyle\Delta m_{ij}^{2}(t)=m_{i}^{2}-m_{j}^{2}\approx\Delta m^{2}_{ij,o}[1+2\epsilon_{\phi}\cos{(m_{\phi}t)}] (6)

where Δmij,o2\Delta m^{2}_{ij,o} is the standard (undistorted) value and Δm2(t)\Delta m^{2}(t) evolves through ϕ(t)\phi(t) (see Equation 3), with an amplitude ϵϕ\epsilon_{\phi} (see Equation 5), and a frequency mϕm_{\phi}. The mass-squared difference Δmij2\Delta m^{2}_{ij} between neutrinos of different flavors follows the usual convection (e.g., Lopes, 2017) such that Δmi12=mi2m12\Delta m^{2}_{i1}=m^{2}_{i}-m^{2}_{1} (i=i=2, 33, 44). In particular for the sterile neutrino, we have Δm412=m42m12\Delta m^{2}_{41}=m^{2}_{4}-m^{2}_{1} where m4m_{4} is the mass of the sterile neutrino. Similarly, the mixing angles variation is written as

θij(t)θij,o+ϵϕcos(mϕt),\displaystyle\theta_{ij}(t)\approx\theta_{ij,o}+\epsilon_{\phi}\cos{(m_{\phi}t)}, (7)

where θij,o\theta_{ij,o} is the standard (undistorted) mixing angle. The indexes ii and jj in θij\theta_{ij} follow a convention identical but not equal for the mass-squared differences (see Lopes , 2018a, and references therein). Therefore, as first suggested by Krnjaic et al. (2018), the LDM ϕ(t)\phi(t) impacts the neutrino flavor oscillations through the modified expressions for the mass-squared differences (Equation 6) and mixing angles (Equation 7).

3. Light Dark Matter and the Sterile Neutrino Model

In the following section, we consider a 3+1 neutrino flavor oscillation model to describe the propagation of active neutrinos (νe\nu_{e}, ντ\nu_{\tau}, νμ\nu_{\mu}) plus a sterile neutrino νs\nu_{s} through the solar plasma. Following the usual notation (νe,ντ,νμ,νs)(\nu_{e},\nu_{\tau},\nu_{\mu},\nu_{s}) corresponds to the neutrino flavors, (ν1\nu_{1}, ν2\nu_{2}, ν3\nu_{3}, ν4\nu_{4}) are the mass neutrino eigenstates, and (m1m_{1}, m2m_{2}, m3m_{3}, m4m_{4}) are the neutrino masses. The evolution of neutrinos propagating in matter is described by the see

idΨdr=Ψ=12E(𝐔M2𝐔+2E𝒱)Ψ,\displaystyle i\frac{d\Psi}{dr}={\cal H}\Psi=\frac{1}{2E}\left(\mathbf{U}M^{2}\mathbf{U}^{\dagger}+2E{\cal V}\right)\Psi, (8)

where {\cal H} is the Hamiltonian and Ψ=(νe,ντ,νμ,νs)T\Psi=(\nu_{e},\nu_{\tau},\nu_{\mu},\nu_{s})^{T}. M2M^{2} is a neutrino mass matrix, 𝐔\mathbf{U} is a (4×44\times 4) unitary matrix describing the mixing of neutrinos and 𝒱{\cal V} is the diagonal matrix of Wolfenstein potentials (Kuo and Pantaleone, 1989). M2M^{2} is defined as M2=diag{0,Δm212,Δm312,Δm412}M^{2}=\mathrm{{diag}}\{0,\Delta m^{2}_{21},\Delta m^{2}_{31},\Delta m^{2}_{41}\}. The first term of the Hamiltonian describes the neutrino propagation through vacuum and the second term incorporates the matter effects or Mikheyev-Smirnov-Wolfenstein (MSW) effects (Wolfenstein, 1978; Mikheyev and Smirnov, 1985). In general, the Hamiltonian {\cal H} that drives the evolution of neutrino flavor must include the Wolfenstein potentials related with ϕ(t)\phi(t) (Brdar et al., 2018).

In most studies of three-neutrino flavor models, the authors are solely interested in the modulation coming from the square mass differences Δmij2(t)\Delta m^{2}_{ij}(t) (by Equation 6) and mixing angles θij(t)\theta_{ij}(t) (by Equation 7). For that reason, all neutrinos are assumed to couple ϕ(t)\phi(t). As a consequence, their contribution to 𝒱{\cal V} cancels out. Hence, it is correct to neglect the contribution of ϕ(t)\phi(t) to the Wolfenstein potential (Dev et al., 2020). Nevertheless, here in this 3+1 neutrino flavor model, as we will discuss later, we include the contribution of ϕ(t)\phi(t) in 𝒱{\cal V}.

This 3+1 neutrino flavor model with dark matter is identical to the standard (undistorted) three-neutrino flavor model (see Equation 8). However, in this model we included a sterile neutrino, and the Wolfenstein potentials in 𝒱{\cal V} are modified to take into account the new LDM field ϕ\phi (Miranda et al., 2015).

3.1. Neutrino Matter-induced Oscillations

Refer to caption
Figure 1.— The survival probability electron neutrinos Pee(E,ϕ)P_{ee}(E,\phi) and Pee(E)\langle P_{ee}(E)\rangle in a 3+1 neutrino flavor oscillation model in which neutrino couple to the light dark matter field ϕ\phi with fixed value of mϕm_{\phi} and ϵϕ\epsilon_{\phi} (see main text). The figure shows Pee(E)\langle P_{ee}(E)\rangle (with ϵϕ0%\epsilon_{\phi}\approx 0\%) for the following LDM models with Gϕ/mϕG_{\phi}/m_{\phi} ratios: 1029GFeV110^{29}\;{\rm G_{\rm F}\,eV^{-1}} (blue curve), 1030GFeV110^{30}\;{\rm G_{\rm F}\,eV^{-1}} (green curve), 1031GFeV110^{31}\;{\rm G_{\rm F}\,eV^{-1}} (orange curve). The figure also shows a model with Gϕ0G_{\phi}\approx 0 (red curve) and an ensemble of Pee(E,ϕ)P_{ee}(E,\phi) corresponding to different time-varying mass-square differences and angles for ϵϕ=1.5%\epsilon_{\phi}=1.5\% (pink band). See main text for details.

In the standard three-neutrino flavor model111In this model, the intermediate particle is an heavy boson, specifically the ZZ or W±W^{\pm} bosons., the matter potential 𝒱{\cal V} takes into account the interaction of active neutrinos (νe,νμ,ντ\nu_{e},\nu_{\mu},\nu_{\tau}) with the ordinary fermions of the solar plasma, for which the 𝒱=diag{Vcc+Vnc,Vnc,Vnc}{\cal V}=\mathrm{{diag}}\{V_{cc}+V_{\rm nc},V_{\rm nc},V_{\rm nc}\} where VccV_{\rm cc} corresponds to the weak charged current (cc\rm cc) that takes into account the forward scattering of νe\nu_{e} with electrons, and VncV_{\rm nc} is the weak neutral current (nc\rm nc) that corresponds to the scattering of the active neutrinos with the ordinary fermions of the solar plasma (e.g., Xing, 2020). VncV_{\rm nc} can be expressed as Vnc=Vnce+Vncp+VncnV_{\rm nc}=V^{e}_{\rm nc}+V^{p}_{\rm nc}+V^{n}_{\rm nc} where VncjV^{j}_{\rm nc} with j=j= ee, pp, nn are the contributions coming from electrons, protons, and neutrons, respectively. However due to the electrical neutrality of the solar plasma, the contribution of VnceV^{e}_{\rm nc} and VncpV^{p}_{\rm nc} canceled out such that Vnc=VncnV_{\rm nc}=V^{n}_{\rm nc}. Accordingly, Vcc=2GFne(r)V_{\rm cc}=\sqrt{2}G_{\rm F}\;n_{e}(r) and Vnc=Vncn=GF/2nn(r)V_{\rm nc}=V^{n}_{\rm nc}=G_{\rm F}/\sqrt{2}\;n_{n}(r). Here GFG_{\rm F} is the Fermi constant and ne(r)n_{e}(r) and nn(r)n_{n}(r) are the number density of electrons and neutrons inside the Sun. Nevertheless, since VncV_{\rm nc} is an universal term for all active neutrino flavors, and as such does not change the flavor oscillations pattern, conveniently we write 𝒱=diag{Vcc+0,0,0}{\cal V}=\mathrm{{diag}}\{V_{\rm cc}+0,0,0\}. Now, the inclusion of sterile neutrinos in the neutrino flavor model alters 𝒱{\cal V} (from Equation 8) by incorporating a new degree of freedom, as a consequence 𝒱=diag{Vcc+Vnc,Vnc,Vnc,0}{\cal V}=\mathrm{{diag}}\{V_{\rm cc}+V_{\rm nc},V_{\rm nc},V_{\rm nc},0\} (e.g., Giunti and Li, 2009; Maltoni and Smirnov, 2016; Xing, 2020).

Finally, in our 3+1 neutrino flavor model, we include the interaction of active and sterile neutrinos with the dark matter field ϕ\phi by means of an intermediate heavy boson II222 We assume the boson II has a mass mIm_{I} identical to ZZ and W±W^{\pm} bosons.. These interactions result from the forward scattering of these neutrinos through the LDM field ϕ\phi, thus 𝒱=diag{Vcc+Vnc+Vνeϕ,Vnc+Vνμϕ,Vnc+Vντϕ,Vnc+Vνsϕ}{\cal V}=\mathrm{{diag}}\{V_{\rm cc}+V_{\rm nc}+V_{\nu_{e}\phi},V_{\rm nc}+V_{\nu_{\mu}\phi},V_{\rm nc}+V_{\nu_{\tau}\phi},V_{\rm nc}+V_{\nu_{s}\phi}\}, where VνiϕV_{\nu_{i}\phi} (with νi=νe,νμ,ντ,νs\nu_{i}=\nu_{e},\nu_{\mu},\nu_{\tau},\nu_{s}) relates to the neutrino νi\nu_{i}. This 𝒱{\cal V} corresponds to a generalization of the Wolfenstein potentials found in the literature, for which most neutrino flavor models only take into account the scattering of the sterile neutrinos on heavy dark matter (Capozzi et al., 2017; Lopes , 2018a; Lopes and Silk, 2019).

In our model, we opt to assume that all active neutrinos experience the same interaction with the LDM field ϕ\phi, such that their dark matter potentials are the same, such that Vνjϕ=VνaϕV_{\nu_{j}\phi}=V_{\nu_{a}\phi} (with j=e,μ,τj=e,\mu,\tau), it follows 𝒱=diag{Vcc+Vnc+Vνaϕ,Vnc+Vνaϕ,Vnc+Vνaϕ,Vnc+Vνsϕ}{\cal V}=\mathrm{{diag}}\{V_{\rm cc}+V_{\rm nc}+V_{\nu_{a}\phi},V_{\rm nc}+V_{\nu_{a}\phi},V_{\rm nc}+V_{\nu_{a}\phi},V_{\rm nc}+V_{\nu_{s}\phi}\}. Now, if we subtract the common term Vnc+VνaϕV_{\rm nc}+V_{\nu_{a}\phi} to the diagonal matrix 𝒱{\cal V}, the latter takes the simple form: 𝒱=diag{Vcc,0,0,VνsϕVνaϕVnc}{\cal V}=\mathrm{{diag}}\{V_{\rm cc},0,0,V_{\nu_{s}\phi}-V_{\nu_{a}\phi}-V_{\rm nc}\}.

The potential VνiϕV_{\nu_{i}\phi} (with i=a,si=a,s) is given by Vνiϕ=GνiϕnϕV_{\nu_{i}\phi}=G_{\nu_{i}\phi}n_{\phi} where GνiϕG_{\nu_{i}\phi} is the equivalent of the Fermi constant and nϕn_{\phi} is the distribution of dark matter inside the Sun (Smirnov and Xu, 2019). Equally, VνiϕV_{\nu_{i}\phi} relates directly with the local density of dark matter ρDM\rho^{\odot}_{\rm DM} by the expression: Vνsϕ=(Gνsϕ/mϕ)(ρDMΩϕ/ΩDM)V_{\nu_{s}\phi}=\left(G_{\nu_{s}\phi}/m_{\phi}\right)\left(\rho^{\odot}_{\rm DM}\Omega_{\phi}/\Omega_{\rm DM}\right), where we assume the ratio Gνsϕ/mϕG_{\nu_{s}\phi}/m_{\phi} is a free parameter of the LDM model. The generalized Fermi constant is defined as Gνiϕ=gνigϕ/mI2G_{\nu_{i}\phi}=g_{\nu_{i}}g_{\phi}/m_{I}^{2} where gνig_{\nu_{i}} represents the coupling constant of the corresponding neutrino νi\nu_{i}, and mIm_{I} is the mass of the intermediate boson II (Miranda et al., 2015). This expression for the potential VνiϕV_{\nu_{i}\phi} is valid since we assume that mI1Rm_{I}^{-1}\ll R_{\odot} where RR_{\odot} is the solar radius333As an example, if we consider mImZ90GeVm_{I}\approx m_{Z}\approx 90GeV where mZm_{Z} is the mass of the ZZ boson, then the propagation of neutrinos (like of the neutral current) verifies the condition mI1Rm_{I}^{-1}\ll R_{\odot}. (Smirnov and Xu, 2019). In general, we could expect that the contribution of ϕ(t)\phi(t) to VνiϕV_{\nu_{i}\phi} could lead to a time-dependent relation, however, as discussed previously (in section 2.1) and mentioned for the first time by Khmelnitsky and Rubakov (2014), this is because the oscillatory component on the local density relates with vϕ2v_{\phi}^{2}. This term is minimal, and therefore we neglected it.

In this preliminary study, without loss of generality, we choose to simplify 𝒱{\cal V} further: since the term VνsϕVνaϕVncV_{\nu_{s}\phi}-V_{\nu_{a}\phi}-V_{\rm nc} has two Wolfenstein potentials ( VνaϕV_{\nu_{a}\phi} and VνsϕV_{\nu_{s}\phi}) that effectively correspond to two new degrees of freedom, both of these have an identical impact on the neutrino flavor oscillation model. We choose to simplify the model by assuming that VνaϕV_{\nu_{a}\phi} is much smaller than VncV_{\rm nc}. Consequently, 𝒱{\cal V} takes the simplified form: 𝒱diag{Vcc,0,0,VνsϕVnc}{\cal V}\approx\mathrm{{diag}}\{V_{\rm cc},0,0,V_{\nu_{s}\phi}-V_{\rm nc}\}. For reference, we note that in the Sun’s core VncV_{\rm nc} is always smaller than VccV_{\rm cc}, once nen_{e} is more than twice as lager as nnn_{n} (e.g. Lopes, 2018b). This potential is identical to others found in the literature, for instance in Capozzi et al. (2017) and Lopes (2018a). Therefore, the matter potential VνsϕV_{\nu_{s}\phi} reads Vνsϕ=GνsϕnϕV_{\nu_{s}\phi}=G_{\nu_{s}\phi}n_{\phi} where for convenience of analysis, we choose to define the generalized Fermi constant as Gνsϕ=42GϕGFG_{\nu_{s}\phi}=4\sqrt{2}G_{\phi}G_{\rm F} where GϕG_{\phi} is our free parameter. Since these dark matter particles have a mass much smaller than 4GeV4\;{\rm GeV}, the solar plasma conditions do not allow the accretion of dark matter by the Sun (e.g., Lopes and Lopes, 2019), therefore we will assume that the distribution of dark matter inside the star is equal to the value measured for the solar neighborhood nϕn_{\phi} (see section 2.1).

Refer to caption
Figure 2.— The survival probability electron neutrinos Pee(E,ϕ)P_{ee}(E,\phi) and the corresponding Pee(E)\langle P_{ee}(E)\rangle for some of the models shown in figure 1 for which ϵϕ=1.5%\epsilon_{\phi}=1.5\% : Gϕ=1020GFG_{\phi}=10^{20}\;G_{\rm F} (blue band and dark blue curve), Gϕ=1021GFG_{\phi}=10^{21}\;G_{\rm F} (green band and dark green curve) and Gϕ=1022GFG_{\phi}=10^{22}\;G_{\rm F} (orange band and dark red curve). The figure also shows Pee(E)\langle P_{ee}(E)\rangle for the 3+1 model with no time dependence (Gχ0G_{\chi}\approx 0) and ϵϕ0\epsilon_{\phi}\approx 0) as a thin dashed black curve. The latter curve corresponds to the red curve in figure 1.

3.2. Neutrino Flavor Oscillation Model and the Survival Probability of Electron Neutrinos

If we adopt as reference the current experimental set of parameters for the active neutrinos (e.g., Esteban et al., 2019), the propagation neutrinos in the solar interior are completely adiabatic. The same is valid for the 3+1 neutrino flavor oscillation model coupled to an LDM field ϕ\phi considered in this study. Conveniently, the propagation of neutrinos away from resonances is well represented by a two neutrino flavor oscillation model. The motivation for such approximation can be found in Lopes (2018b) and references therein. In such a case, the electron neutrino flavor oscillation is dominated by the (ν1,ν2)(\nu_{1},\nu_{2}) mass eigenstates and is only slightly affected by the decoupled (ν3,ν4)(\nu_{3},\nu_{4}) eigenstates, since the associated mixing angles for the latter pair are very small (Kuo and Pantaleone, 1986). Moreover, ν3\nu_{3} and ν4\nu_{4} evolve independent of each other and are completely independent of the doublet (ν1,ν2)(\nu_{1},\nu_{2}). In this limit, as proposed by several authors (e.g, Palazzo, 2011; Blennow and Smirnov, 2013), the split of the 3+1 neutrino flavor model into a dominant two neutrino flavor model (νe,νμ\nu_{e},\nu_{\mu}) with additional corrections for ντ\nu_{\tau} and νs\nu_{s} significantly simplified the calculation and allowed us to obtain an analytical solution (e.g., Kuo and Pantaleone, 1989).

Among the many expressions available in the literature to compute the survival probability of electron neutrinos PeP_{e} (e.g., Lunardini and Smirnov, 2000; Miranda et al., 2015) in a 3+1 neutrino flavor model developed in the approximate scenario of a two-flavor neutrino model (e.g., Kuo and Pantaleone, 1989), we opted to choose the expression obtained by Capozzi et al. (2017) for the case in which VccE/Δm3121V_{\rm cc}E/\Delta m_{31}^{2}\ll 1 (and s34=0s_{34}=0) which has a better numerical accuracy than others. In that case the survival probability of electron neutrinos, i.e., PeP_{e} [P(νeνe]][\equiv P(\nu_{e}\rightarrow\nu_{e}]], reads

Pee(E,ϕ)=s134+c134c244c144+am+bm,\displaystyle P_{ee}(E,\phi)=s_{13}^{4}+c_{13}^{4}c_{24}^{4}c_{14}^{4}+a_{m}+b_{m}, (9)

where cij=cosθijc_{ij}=\cos{\theta_{ij}} and sij=sinθijs_{ij}=\sin{\theta_{ij}}. The functions ama_{m} and bmb_{m} are dependent on the internal structure of the Sun and are given by the expressions:

am=C1(smc14cms14s24)2\displaystyle a_{m}=C_{1}(s_{m}c_{14}-c_{m}s_{14}s_{24})^{2} (10)

and

bm=C2(cmc14+sms14s24)2\displaystyle b_{m}=C_{2}(c_{m}c_{14}+s_{m}s_{14}s_{24})^{2} (11)

where cm=cosθmc_{m}=\cos{\theta_{m}}, sm=sinθms_{m}=\sin{\theta_{m}}, C1=c134(c14s12c12s14s24)2C_{1}=c_{13}^{4}(c_{14}s_{12}-c_{12}s_{14}s_{24})^{2} and C2=c134(c12c14+s12s14s24)2C_{2}=c_{13}^{4}(c_{12}c_{14}+s_{12}s_{14}s_{24})^{2}. The angle θm\theta_{m} is obtained for the present-day Sun (i.e., the standard solar model, see details of this model in Lopes and Silk, 2013) using the expression (Capozzi et al., 2017): cos(2θm)=x(y2+x2)1/2,\cos{(2\theta_{m})}={\cal M}_{x}({\cal M}_{y}^{2}+{\cal M}_{x}^{2})^{-1/2}, where xcos(2θ12)ηνVx{\cal M}_{x}\equiv\cos{(2\theta_{12})}-\eta_{\nu}V_{x} and y|sin(2θ12)+ηνVy|{\cal M}_{y}\equiv|\sin{(2\theta_{12})}+\eta_{\nu}V_{y}|. ην\eta_{\nu} is the ratio of the energy of the neutrino EE in relation to Δm212\Delta m^{2}_{21} given by ην(E)=4E/Δm212\eta_{\nu}(E)=4E/\Delta m_{21}^{2}. The functions VxV_{x} and VyV_{y} are given by

Vx=12[Vccc132(c142s142s242)+Vs(s142c142s242)],\displaystyle V_{x}=\frac{1}{2}\left[V_{\rm cc}c^{2}_{13}(c^{2}_{14}-s_{14}^{2}s_{24}^{2})+V_{s}(s_{14}^{2}-c_{14}^{2}s_{24}^{2})\right], (12)
Vy=(VsVccc132)c14s14s24\displaystyle V_{y}=(V_{s}-V_{\rm cc}c_{13}^{2})c_{14}s_{14}s_{24} (13)

and

Vs=VνsϕVnc.\displaystyle V_{s}=V_{\nu_{s}\phi}-V_{\rm nc}. (14)
Refer to caption
Figure 3.— The survival probability of the electron neutrinos for several solar nuclear reactions: pppp (yellow diamond), B7e{}^{7}Be (green square), 7Be (red upward triangle), peppep (blue downward triangle) and 8B HER (salmon circle), 8B HER-I (orange circle), B8{}^{8}B HER-II (magenta circle) and B8{}^{8}B (cyan square). The diamond, triangle, and circles points use data from the Borexino (Borexino Collaboration et al. , 2020; Agostini et al., 2019; Borexino Collaboration et al., 2018; Bellini et al., 2010) and the two square data points correspond to data obtained from SNO (Aharmim et al., 2013) and Super-Kamiokande (Abe et al., 2016; Cravens et al., 2008) detectors. Each color band corresponds to an ensemble of Pee(E,ϕ)P_{ee}(E,\phi) (Equation 9) in a 3+1 neutrino model with Gϕ0G_{\phi}\approx 0 and ϵϕ0\epsilon_{\phi}\approx 0 (dashed curve), 1.5% (green), 3.0% (orange) and 6.0% (light blue). These survival probabilities are computed using an high-Z SSM (see the main text) and the experimental values from the different neutrino experiments.

3.3. Light Dark Matter Impact on Solar Neutrinos

The survival probability of electron neutrinos (Equation 9) is a time-dependent function through equations (6), (7) and (3). Conveniently we define an effective oscillation probability Pee(E)\langle P_{ee}(E)\rangle that corresponds to an ensemble average of all the Pee,(E,ϕ)P_{ee,}(E,\phi) (Equation 9), as such

Pee(E)=0τϕPee(E,ϕ)dtτϕ,\displaystyle\langle P_{ee}(E)\rangle=\int_{0}^{\tau_{\phi}}P_{ee}(E,\phi)\frac{dt}{\tau_{\phi}}, (15)

where τϕ=2π/mϕ\tau_{\phi}=2\pi/m_{\phi} is the period of the LDM field ϕ(t)\phi(t).

The ability of a solar neutrino detector to measure the impact of the time-dependent LDM field ϕ(t)\phi(t) on the survival probability Pee(E,ϕ)P_{ee}(E,\phi) (Equation 9) depends on three characteristic time scales: the neutrino flight time τν\tau_{\nu}, the time between two consecutive neutrino detections τev\tau_{ev}, and the total run time of the experiment τex\tau_{ex}. The neutrino flight time is proportional to the Earth-Sun distance dd_{\earth} such that τν=d/c8.2min\tau_{\nu}=d_{\earth}/c\approx 8.2\;{\rm min} where cc is the speed of light. The number of events measured by a detector varies strongly from one to another.

The next generation of experiments will have τev\tau_{ev} much larger than the pioneer Homestake experiment that only detects a few events per year (Bahcall and Davis, 1976). The forthcoming Jiangmen Underground Neutrino Observatory (JUNO; Adam et al., 2015) experiment expects to measure a few tens of neutrinos per day (for instance 200 events per day or τev7min\tau_{ev}\approx{\rm 7\;min}). The total experimental run time for most solar neutrino detectors is of the order of a few decades (for instance τex10yr\tau_{ex}\approx{\rm 10\;yr}), and future experiments will also have significant running times. Hence for all models considered in this study, we assume that solar neutrino detectors will run for long periods and will collect a large number of events, therefore we assume that τev\tau_{ev} and τex\tau_{ex} have sufficient small and large values, respectively.

In such conditions, the solar neutrino spectra time modulation by ϕ(t)\phi(t) depends on the period τϕ\tau_{\phi} of the LDM field in comparison to the flight time of solar neutrinos τν\tau_{\nu}. Since these neutrinos have a τν8.2min\tau_{\nu}\approx 8.2\;{\rm min}, it is possible to find the value of mϕm_{\phi} for which τν=τϕ\tau_{\nu}=\tau_{\phi} which occurs for mϕ,c=8.3 1018eVm_{\phi,c}=8.3\;10^{-18}\;{\rm eV}. Accordingly, we can define two regimes for the time modulation of survival probability of electron neutrinos :

  1. 1.

    For τϕτν\tau_{\phi}\geq\tau_{\nu} (low-frequency regime or low LDM mass), the time modulation of Pee(E,ϕ)P_{ee}(E,\phi) occurs when the period of ϕ(t)\phi(t) is larger than τν\tau_{\nu}. In this case a temporal variation of the neutrino signal may be observed. This corresponds to a LDM field with a mass such that mϕmϕ,cm_{\phi}\leq m_{\phi,c}. Therefore, the LDM field can induce an observable time variation in neutrino oscillation measurements as periodicity in the solar neutrino fluxes (Berlin, 2016). Obviously, if τϕ\tau_{\phi} becomes very large, the modulation of Pee(E,ϕ)P_{ee}(E,\phi) becomes indistinguishable from the standard scenario (undistorted case), since the running time of the experiment is not sufficient to observe this phenomena. Nevertheless, in our study, the LDM field has always an mϕ1023eVm_{\phi}\geq 10^{-23}\;{\rm eV} or a period τϕ13yr\tau_{\phi}\geq 13\;{\rm yr}. Therefore, it is always possible to probe such a model with current experimental running times.

  2. 2.

    For τϕτν\tau_{\phi}\leq\tau_{\nu} (high-frequency regime or high LDM mass), the change of Pee(E,ϕ)P_{ee}(E,\phi) due to ϕ(t)\phi(t) is too fast to be observed as a modulating signal like in the previous case. This regime occurs for LDM fields with a mass such that mϕmϕ,cm_{\phi}\geq m_{\phi,c}. Nevertheless, the time average of the ensemble of oscillation probability Pee(E,ϕ)P_{ee}(E,\phi) can be distorted in such a regime, hence the effect can be detected as Pee(E)\langle P_{ee}(E)\rangle which will deviate from the standard scenario (Krnjaic et al., 2018). The net effect of averaging over time induces a shift in the observed values of Pee(E,ϕ)P_{ee}(E,\phi) relative to its undistorted value.

Therefore, we can expect to study both regimes in a quite reasonable range of LDM masses using data from the present and future solar neutrino experiments. In fact, some of the current solar neutrino detectors have already large statistics and high event rates that we can use to look for time modulations in solar neutrinos. Some of these neutrino collaborations have already searched for regular phenomena with periods varying from 10 minutes to 10 yr (e.g., Yoo et al., 2003; Aharmim et al., 2010).

In this work, we will study models that will fall in these two regimes of time modulation. Therefore to satisfy the conditions mentioned above, we decided to analyze the impact of the LDM field in solar neutrino fluxes for ϕ(t)\phi(t) with a period τϕ\tau_{\phi} varying from 4μs4\;\mu{\rm s} to 13yr13\;{\rm yr} or equivalently with a mϕm_{\phi} varying from 109eV10^{-9}\;{\rm eV} to 1023eV10^{-23}\;{\rm eV}, which is a range possible to be scanned by future detectors like Deep Underground Neutrino Experiment (DUNE Collaboration et al., 2015) and JUNO (An et al., 2016).

Refer to caption
Figure 4.— The survival probability of the electron neutrinos for several LDM models (see the main text and caption of figure 3 and 2 for details). The LDM models are identical to the ones presented in Figure 2, with the following changes: (i) Gϕ=0G_{\phi}=0 and ϵϕ=1.5%\epsilon_{\phi}=1.5\% (red band); (ii) Gϕ=1020GFG_{\phi}=10^{20}G_{\rm F} and ϵϕ=3.0%\epsilon_{\phi}=3.0\% (light-blue band); (iii) Gϕ=1021GFG_{\phi}=10^{21}G_{\rm F} and ϵϕ=1.5%\epsilon_{\phi}=1.5\% (green band); (iv) Gϕ=1022GFG_{\phi}=10^{22}G_{\rm F} and ϵϕ=3.0%\epsilon_{\phi}=3.0\% (orange band). The data points correspond to the same ones displayed in figure 3. However, the data with the lowest energy corresponds to the expected precision to be attained by the Darwin experiment in measuring Pee±ΔPeeP_{ee}\pm\Delta P_{ee}, for which ΔPee\Delta P_{ee} could be as low as ΔPee=0.017\Delta P_{ee}=0.017 (Aalbers et al., 2020). The continuous dashed curve corresponds to a 3+1 neutrino model with Gϕ0G_{\phi}\approx 0 and ϵ=0%\epsilon=0\%.

4. Light Dark Matter Impact on Electron Neutrino Spectra

Inside the Sun, the flux variation of neutrinos with different flavors due to matter (including LDM) is strongly dependent of the local distributions of electrons and neutrons, but also on the population of dark matter particles in the solar neighbourhood. This new flavor mechanism (sterile neutrinos and LDM field ϕ\phi) affects all electron neutrinos produced in the Sun’s core. A detailed discussion about the neutrino sources inside the Sun, and their specific solar properties, can be found in Lopes (2013, 2017). The average survival probability of electron neutrinos for each nuclear reaction in the solar interior, i.e., Pe,i(E,ϕ)P_{e,i}(E,\phi) is computed by

Pee,i(E,ϕ)=Ci0RPe(E,ϕ,r)Si(r)4πρ(r)r2𝑑r,\displaystyle P_{ee,i}(E,\phi)=C_{i}\int_{0}^{R_{\odot}}P_{e}(E,\phi,r)S_{i}(r)4\pi\rho(r)r^{2}dr, (16)

where CiC_{i} (=[0RSi(r)4πρ(r)r2𝑑r]1)\left(=\left[\int_{0}^{R_{\odot}}S_{i}(r)4\pi\rho(r)r^{2}\;dr\right]^{-1}\right) is a normalization constant and Si(r)S_{i}(r) is the electron neutrino emission function for the ii solar nuclear reaction. ii corresponds to the following solar neutrino sources (from the PP chain and CNO cycle nuclear reactions): pppp, peppep, 8B, 7Be, 13N, 15O and 17F.

Moreover, since the survival probabilities Pee,i(E,ϕ)P_{ee,i}(E,\phi) (Equation 16) are time dependent through ϕ\phi, these quantities also vary with time. Therefore, the oscillation probability Pee(E)\langle P_{ee}(E)\rangle (Equation 15) is generalized for each specific nuclear reaction ii:

Pee,i(E)=0τϕPe,i(E,ϕ)dtτϕ.\displaystyle\langle P_{ee,i}(E)\rangle=\int_{0}^{\tau_{\phi}}P_{e,i}(E,\phi)\frac{dt}{\tau_{\phi}}. (17)

The LDM field ϕ\phi can lead to different temporal imprints on the neutrino oscillation measurements. The specific impact depends on the mass of the LDM particle. In the following, we compute the spectra of neutrinos from any specific nuclear reaction that we know to be essentially independent of the properties surrounding solar plasma. Since in the 3+1 neutrino flavor model new processes exist to change the survival probability of electron neutrinos, this will modify the solar neutrino spectra measured on Earth. These new processes will alter the conversion rates of νe\nu_{e} to other flavors (νμ\nu_{\mu}, ντ\nu_{\tau} and νs\nu_{s}) and vice versa. Accordingly, the electron neutrino spectrum of the nuclear reaction ii inside the core is defined as Φi\Phi_{i}, and Φi\Phi_{i\odot} is the electron neutrino spectrum arriving on Earth (Lopes, 2018b) such that:

Φi,(E)=Pee,i(E,ϕ)Φi(E)\displaystyle\Phi_{i,\odot}(E)=P_{ee,i}(E,\phi)\Phi_{i}(E) (18)

where Pee,i(E)P_{ee,i}(E) is the average survival probability of electron neutrinos for reactions in the solar interior as given by equation 16. Equally if we take the time average of equation 18, we obtain the following averaged spectrum for each nuclear reaction ii:

Φi,(E)=Pee,i(E)Φi(E),\displaystyle\Phi_{i,\odot}(E)=\langle P_{ee,i}(E)\rangle\Phi_{i}(E), (19)

where Pee,i(E)\langle P_{ee,i}(E)\rangle is the average survival probability of electron neutrinos as given by equation 17.

Refer to caption
Figure 5.— The B8{}^{8}B solar electron neutrino spectra measured on Earth detectors. The blue curve corresponds to the standard (undistorted) neutrino spectra, the cyan curve to the averaged (distorted) electron spectra, and the red band corresponds to the ensemble of spectra related with the time-dependent flavor parameters. This LDM model has a Gϕ0G_{\phi}\approx 0 and an ϵϕ=3.0%\epsilon_{\phi}=3.0\%. In the calculation of the 8B spectrum we use a high-Z SSM (see the main text).
Refer to caption
Figure 6.— The pppp solar electron neutrino spectra measured on Earth detectors. The LDM model and color scheme is the same one as in Figure 5. The error bar curves (yellow curves) on the averaged (distorted) electron spectra (cyan curve) corresponds to the expected precision to be attained by the Darwin experiment (Aalbers et al., 2020). See Figure 4 and main the text.

5. The Sun: Light Dark Matter and Sterile Neutrinos

Here, we will study the impact of the theoretical model presented in the previous sections, specifically we compute the survival probability of electron neutrinos (as given by equations (9), (15), (16) and (17)) in the case of a standard solar model with low-Z (e.g., Lopes and Silk, 2013; Capelo and Lopes, 2020).

In the parameterization for the 3+1 neutrino flavor oscillation model, we opt to adopt the recent values obtained in the data analysis of the standard three-neutrino flavor oscillation model obtained by de Salas et al. (2020), and for the sterile neutrino additional fiducial parameters we used the values obtained by Gariazzo et al. (2015). Accordingly, for a parameterisation with a normal ordering of neutrino masses, the mass-square difference and the mixing angles have the following values: Δm212=7.500.20+0.22×105eV2\Delta m^{2}_{21}=7.50^{+0.22}_{-0.20}\times 10^{-5}{\rm eV^{2}}, sin2θ12=0.318±0.016\sin^{2}{\theta_{12}}=0.318\pm 0.016, and sin2θ13=0.022500.00078+0.00055\sin^{2}{\theta_{13}}=0.02250^{+0.00055}_{-0.00078}. Although, Δm312=2.560.0004+0.0003×103eV2\Delta m^{2}_{31}=2.56^{+0.0003}_{-0.0004}\times 10^{-3}{\rm eV^{2}} and sin2θ23=0.560.022+0.016\sin^{2}{\theta_{23}}=0.56^{+0.016}_{-0.022}, we mention them here for reference (de Salas et al., 2020). These new parameters are consistent with previous estimations (Esteban et al., 2019; Gonzalez-Garcia et al., 2016). For the sterile neutrino, we choose the following fiducial values for the mass-square difference and mixing angles  (Gariazzo et al., 2015, 2016; Capozzi et al., 2017): Δm412=1.6eV2\Delta m^{2}_{41}=1.6\ {\rm eV^{2}}, sin2θ14=0.027\sin^{2}{\theta_{14}}=0.027, sin2θ42=0.014\sin^{2}{\theta_{42}}=0.014 and the other mixing angle for the sterile neutrinos are fixed to zero. Moreover, we assume that all phases (δ13,14,34\delta_{13,14,34}) and other angles related to the sterile neutrino are equal to zero.

The present-day internal structure of the Sun corresponds to an up-to-date standard solar model (SSM) that has a better agreement with neutrino fluxes and helioseismic data sets. This solar model was obtained from a one-dimensional stellar evolution code allowed to evolve in time until the present-day solar age, 4.574.57 Gyr, having been calibrated to the values of luminosity and effective temperature of the present Sun, of 3.8418×10333.8418\times 10^{33} erg s-1 and 57775777 K, respectively, as well as the observed abundance ratio at the Sun’s surface: (Zs/XsZ_{\text{s}}/X_{\text{s}})=0.0181{}_{\odot}=0.0181, where ZsZ_{s} and XsX_{s} are the metal and hydrogen abundances at the surface of the star (Turck-Chieze and Lopes, 1993; Bahcall et al., 1995, 2006). This stellar model was computed with the release version 12115 of the stellar evolution code MESA (Paxton et al., 2011, 2019). The details about the physics of this standard solar model in which we use the AGSS09 (low-Z) solar abundances (Asplund et al., 2009) are described in Lopes and Silk (2013) and Capelo and Lopes (2020).

Figures 1 and 2 show the impact of the time-dependent mass-square difference (Equation 6) and mixing angles (Equation 7) on the averaged electron survival probability Pee(E)\langle P_{ee}(E)\rangle (Equation 15) for which the LDM field ϕ\phi has a fixed amplitude (Equation 5): ϵϕ=0\epsilon_{\phi}=0 or ϵϕ=1.5%\epsilon_{\phi}=1.5\%. We also show LDM models for which the sterile neutrino couples to ϕ\phi with strength GϕG_{\phi}.

The overall shape of the curve Pee(E)\langle P_{ee}(E)\rangle depends on GϕG_{\phi} times nϕn_{\phi} in the potential VνsϕV_{\nu_{s}\phi} or the ratio Gϕ/mϕG_{\phi}/m_{\phi} as previously mentioned. For instance, in a LDM model in which we fix mϕ=109eVm_{\phi}=10^{-9}\;{\rm eV} (or nϕn_{\phi}), an increase of GϕG_{\phi} from 102010^{20} to 1022GF10^{22}\;G_{\rm F} leads Pee(E)\langle P_{ee}(E)\rangle to vary significantly, as shown in Figure 1. As expected this change in Pee(E)\langle P_{ee}(E)\rangle is more pronounced for high energy neutrinos where the MSW effect is more significant. If we choose higher values of mϕm_{\phi} the results will somehow be similar (see Figure 1).

Evidently, for an LDM model in which mϕm_{\phi} decreases by a certain amount (nϕ=ρϕ/mϕn_{\phi}=\rho_{\phi}/m_{\phi}), the constancy of Gϕ/mϕG_{\phi}/m_{\phi} in VνsϕV_{\nu_{s}\phi} implies that GϕG_{\phi} can increase by the same order of magnitude to obtain the same MSW effect on the Pee(E)\langle P_{ee}(E)\rangle curve (see Figure 1). For instance, an LDM model with mϕ=1023eVm_{\phi}=10^{-23}\;{\rm eV} and Gϕ=107GFG_{\phi}=10^{7}G_{\rm F} will have Pee(E)\langle P_{ee}(E)\rangle identical to an LDM model with mϕ=109eVm_{\phi}=10^{-9}\;{\rm eV} and Gϕ=1021GFG_{\phi}=10^{21}G_{\rm F} , since in both LDM models we have the same ratio: Gϕ/mϕ1030G_{\phi}/m_{\phi}\approx 10^{30}. The same argument explains the reason why the coupling constant between sterile neutrinos and more massive dark matter particles is much smaller in those models than in the present study. For instance, this is the case for particles captured from the dark matter halo by the Sun. Since over time, the star accreted a significant amount of dark matter (e.g., Lopes , 2018a), for these models GϕG_{\phi} is significantly smaller than the value found for the present study.

The most important feature of such a class of LDM models is the time dependence of the dark matter field ϕ(t)\phi(t) and its imprint in the flavor oscillation parameters’ mass-square differences (Equation 6) and mixing angles (Equation 7). As predicted by equation 9, there are many Pee(E,ϕ)P_{ee}(E,\phi) with near similar behavior. Figure 1 shows an ensemble of time-dependent Pee(E,ϕ)P_{ee}(E,\phi) as a pink band. The difference between curves relates to the dependence of the oscillation parameters on time. In this LDM model it is assumed there is a very negligible interaction between sterile neutrinos and ϕ\phi (for which Gϕ0G_{\phi}\approx 0). The figure also shows Pee(E)\langle P_{ee}(E)\rangle (red curve) the time-averaged of the ensemble of Pee(E,ϕ)P_{ee}(E,\phi) curves that we compute using equation 15.

Although there are several parameters that contribute for the time variability of Pee(E,ϕ)P_{ee}(E,\phi) (Equation 9), the main contributions come from θ21\theta_{21} and θ13\theta_{13}. The variability related θ24\theta_{24} is relevant for the high energies. We notice that the contributions coming from Δm212\Delta m^{2}_{21} and θ32\theta_{32} are much smaller than all the parameters mentioned above. The amplitude of the Pee(E,ϕ)P_{ee}(E,\phi) pink band is defined by the value of ϵϕ\epsilon_{\phi} for which we adopt the fiducial value of ϵϕ=1.5%\epsilon_{\phi}=1.5\%. It is worth pointing out that the Pee(E,ϕ)P_{ee}(E,\phi) band is much larger for low-energy than for higher-energy values. Moreover, the averaged value of this ensemble given by Pee(E)\langle P_{ee}(E)\rangle is identical to Pee(E)\langle P_{ee}(E)\rangle with ϵϕ0\epsilon_{\phi}\approx 0. Figure 2 shows the variability of Pee(E,ϕ)P_{ee}(E,\phi) for a LDM with different GϕG_{\phi} values. These results are identical to the model in figure 1. Nevertheless, the LDM model with the largest GϕG_{\phi} has (an orange) band with a smaller amplitude around Pee(E)\langle P_{ee}(E)\rangle. Once again, the band thickness decreases for neutrinos with higher energy for all these models.

Figures 3 and 4 compare our predictions with current solar neutrino data (e.g., Aalbers et al., 2020). These figures show that LDM models with relatively low values of ϵϕ\epsilon_{\phi} and GϕG_{\phi} are compatible overall with current solar neutrino data coming from Borexino, Super-Kamiokande, and SNO. Clearly, this analysis has also shown that the precision of our current solar neutrino experiments is not able to distinguish between some of these LDM models. Nevertheless, it is already possible to put some constraints on these LDM models. For instance, we found that LDM models with Gϕ0G_{\phi}\approx 0 must have a ϵϕ\epsilon_{\phi} smaller than 3%3\% to be consistent with all data, including pppp measurements of the Borexino detector (Borexino Collaboration et al. , 2020) (see Figure 3); and any LDM models must have a ratio Gϕ/mϕG_{\phi}/m_{\phi} smaller than 103010^{30}, otherwise they become inconsistent with pppp and 7Be measurements for several solar detectors (Borexino Collaboration et al. , 2020; Agostini et al., 2019; Borexino Collaboration et al., 2018; Bellini et al., 2010) (see figure 4). Figure 4 shows a LDM model with a mϕ=109eVm_{\phi}=10^{-9}\;{\rm eV} and Gϕ=1022GFG_{\phi}=10^{22}\;G_{\rm F} with a ratio Gϕ/mϕG_{\phi}/m_{\phi} of the order of 1031GF10^{31}\;G_{\rm F} (see Figure 4); This ratio is one order of magnitude larger than the critical Gϕ/mϕG_{\phi}/m_{\phi} value of 103010^{30} discussed in the previous section. Figure 4 also shows that the variability related with time dependence on Pee(E,ϕ)P_{ee}(E,\phi) decreases for large values of GϕG_{\phi}.

There is another important effect that also contributes to the time variability of Pee(E,ϕ)P_{ee}(E,\phi). The PP chain and CNO cycle, nuclear reactions occur at different distances from the center of the Sun and each nuclear reaction emits neutrinos in a well-defined energy range. As a consequence, the electron neutrinos produced in each specific nuclear reaction will be affected differently by the MSW effect. As such, this effect will also contribute to the overall variability of electron neutrinos Pee,i(E,ϕ)P_{ee,i}(E,\phi) (see Equation 16) and their time-averaged Pee,i(E,ϕ)\langle P_{ee,i}(E,\phi)\rangle (see Equation 17).

The time-dependent electron neutrino survival probability will have a significant impact on the neutrino spectra of the different nuclear reactions. Accordingly, figures 5 and 6 show the spectra correspond to two neutrino types: pppp and B8{}^{8}B neutrinos. An essential difference between these two spectra relates to the thickness of the ϵϕ\epsilon_{\phi} band for a fixed value since thickness decreases with neutrino energy. Therefore the ϵϕ\epsilon_{\phi} band is more significant for a pppp spectrum than for a B8{}^{8}B neutrino spectrum. This is an effect identical to the one discussed previously for the Pee(E,ϕ)P_{ee}(E,\phi) functions. Therefore, the measurement of solar neutrino fluxes and solar neutrino spectrum in the energy range below 0.2MeV0.2\;MeV will provide the strongest constraint for such a class of dark matter models. Figure 5 and 6 show the spectra of 8B and pppp, if we assume the precision expected to be attained by the Darwin experiment (Aalbers et al., 2020). Figure 6 also shows the precision expected for the Darwin experiment.

6. Conclusion

This article focuses on the impact of LDM on solar neutrino fluxes, spectra, and survival probabilities of electron neutrinos, specifically a dark matter model made of two particles: a sterile neutrino and an LDM particle. In particular, we describe how the 3+1 neutrino flavor model is affected by this type of LDM particles, with an emphasis on how the LDM affects the Wolfenstein potentials. We also study how the dark matter models affect the survival probability functions of electron neutrinos related to the different nuclear reactions occurring in the solar interior, and we compute the spectra of two relevant solar neutrino sources: pppp and B8{}^{8}B neutrino nuclear reactions.

By studying a large range of dark matter particle masses (from 10910^{-9} to 1023eV10^{-23}\;{\rm eV}) we found that depending on the mass of these LDM particles and the value of the generalized Fermi constant, the shape of electron neutrino survival probability and their spectra can vary with time. We establish that for LDM particles with low masses (low-frequency regime), the solar neutrino detectors can observe the electron neutrino survival probability changing with time. Conversely, for dark matter particles with higher masses (high-frequency regime), this impact can be determined by measuring the time-averaged electron neutrino survival probability.

It was possible to establish, using data of current solar neutrino measurements, that those models with a Gϕ/mϕG_{\phi}/m_{\phi} ratio smaller than 1030GFeV110^{30}\;G_{\rm F}eV^{-1} agree with current solar neutrino data from the Borexino, SNO and Super-Kamiokande detectors. We also found that for models with a near-zero constant, the time-variability amplitude must be smaller than 3%. Such a constraint is equivalent to the condition gϕ2ρDM/(mϕmν)(Ωϕ/ΩDM)1/20.03{g_{\phi}\sqrt{2\rho^{\odot}_{\rm DM}}}/{(m_{\phi}m_{\nu})}\;\left({\Omega_{\phi}}/{\Omega_{\rm DM}}\right)^{1/2}\leq 0.03.

Finally, we also found that the precision expected in the measurements to be made by the Darwin detector will allow us to put powerful constraints to this class of models.

The author is grateful to the MESA team for having made their code publicly available. The author also thanks the anonymous referee for the revision of the manuscript. I.L. thanks the Fundação para a Ciência e Tecnologia (FCT), Portugal, for the financial support to the Center for Astrophysics and Gravitation (CENTRA/IST/ULisboa) through the Grant Project No. UIDB/00099/2020 and grant No. PTDC/FIS-AST/28920/2017.

References