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

Neutrino bound states and bound systems

Alexei Yu. Smirnova and Xun-Jie Xub a Max-Planck-Institut für Kernphysik, Postfach 103980, D-69029 Heidelberg, Germany
b Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, China
(July 17, 2025)
Abstract

Yukawa interactions of neutrinos with a new light scalar boson ϕ\phi can lead to formation of stable bound states and bound systems of many neutrinos (ν\nu-clusters). For allowed values of the coupling yy and the scalar mass mϕm_{\phi}, the bound state of two neutrinos would have the size larger than 101210^{12} cm. Bound states with sub-cm sizes are possible for keV scale sterile neutrinos with coupling y>104y>10^{-4}. For ν\nu-clusters we study in detail the properties of final stable configurations. If there is an efficient cooling mechanism, these configurations are in the state of degenerate Fermi gas. We formulate and solve equations of the density distributions in ν\nu-clusters. In the non-relativistic case, they are reduced to the Lane-Emden equation. We find that (i) stable configurations exist for any number of neutrinos, NN; (ii) there is a maximal central density 109\sim 10^{9} cm-3 determined by the neutrino mass; (iii) for a given mϕm_{\phi} there is a minimal value of Ny3Ny^{3} for which stable configurations can be formed; (iv) for a given strength of interaction, Sϕ=(ymν/mϕ)2S_{\phi}=(ym_{\nu}/m_{\phi})^{2}, the minimal radius of ν\nu-clusters exists. We discuss the formation of ν\nu-clusters from relic neutrino background in the process of expansion and cooling of the Universe. One possibility realized for Sϕ>700S_{\phi}>700 is the development of instabilities in the ν\nu-background at T<mνT<m_{\nu} which leads to its fragmentation. For Sϕ[70, 700]S_{\phi}\in[70,\,700] they might be formed via the growth of initial density perturbations in the ν\nu-background and virialiazation, in analogy with the formation of Dark Matter halos. For allowed values of yy, cooling of ν\nu-clusters due to ϕ\phi-bremsstrahlung and neutrino annihilation is negligible. The sizes of ν\nu-clusters may range from \sim km to 5\sim 5 Mpc. Formation of clusters affects perspectives of detection of relic neutrinos.

I Introduction

In recent years neutrinos and new physics at low energy scales became one of the most active areas of research. In particular, neutrino interactions with new and very light bosons have been actively explored, with considerations covering e.g. oscillation phenomena Samanta (2011); Heeck and Rodejohann (2011); Wise and Zhang (2018); Bustamante and Agarwalla (2019); Smirnov and Xu (2019); Babu et al. (2020) and various astrophysical and cosmological consequences Boehm et al. (2012); Heeck (2014); Huang et al. (2018); Luo et al. (2021); Esteban and Salvado (2021); Green et al. (2021). Yet another possible consequence of such interactions (the subject of this paper) is the formation of bound states and bound systems of neutrinos.

The exploration of possible existence of NN-body bound states of neutrinos (neutrino clouds, balls, stars, or halos) has a long story. In 1964 Markov proposed the existence of “neutrino superstars” formed by massive neutrinos due to the usual gravity Markov (1964). The system is similar to neutron stars and can be described as the degenerate Fermi gas in thermodynamical and mechanical equilibrium. Substituting the mass of neutron, mnm_{n}, by the mass of neutrino, mnmνm_{n}\rightarrow m_{\nu} in results for neutron stars, Markov obtained the radius of a neutrino star:

R8πGN1mν2,R\simeq\sqrt{\frac{8\pi}{G_{N}}}\frac{1}{m_{\nu}^{2}}, (1)

which corresponds to the Oppenheimer-Volkoff limit. In comparison to the radius of neutron star, RR increases by a factor of β2\beta^{2} where

βmnmν=1.91010,\beta\equiv\frac{m_{n}}{m_{\nu}}=1.9\cdot 10^{10},

and for numerical estimation we take mν=0.05m_{\nu}=0.05 eV. Correspondingly, the mass of neutrino star increases and the central density decreases as

Mνβ2MNS,nν(0)β3nNS(0).M_{\nu}\simeq\beta^{2}M_{\rm NS},\,\,\,\,\,n_{\nu}(0)\simeq\beta^{-3}n_{\rm NS}(0). (2)

For mνmem_{\nu}\approx m_{e} used by Markov, the size, the mass and the number density of the superstars would be R=1012R=10^{12} cm, M=106MM=10^{6}M_{\odot} and nν1029cm3n_{\nu}\simeq 10^{29}\,{\rm cm}^{-3} correspondingly. For the present values of neutrino mass (0.1 eV), the relations (1) and (2) give

R51026cm,Mν=41020M,nν(0)108cm3.R\simeq 5\cdot 10^{26}\,{\rm cm},\,\,\,\,\,M_{\nu}=4\cdot 10^{20}M_{\odot},\,\,\,\,\,n_{\nu}(0)\simeq 10^{8}\,{\rm cm}^{-3}.

The radius is only 20 times smaller than the size of observable Universe (the present Hubble scale), and the total number of neutrinos is 10 times larger than the number of relic neutrinos in whole the Universe with the standard density. The bound states of much smaller sizes and higher densities with substantial implications for cosmology and structure formation as well as for laboratory experiments require new physics beyond the Standard Model.

One can also consider heavy (sterile) neutrinos and gravity only. Masses of (10100)(10-100) keV correspond to β=(105104)\beta=(10^{5}-10^{4}), and consequently, stars have radii of (10161014)(10^{16}-10^{14}) cm and masses of (1010108)M(10^{10}-10^{8})M_{\odot}, i.e. much below the galactic scales. The central density would be (10241027)cm3(10^{24}-10^{27})\,{\rm cm}^{-3}. Such a possibility was explored in a series of papers Viollier et al. (1993); Bilic et al. (1999, 2001). This is essentially the warm dark matter for which final configurations of degenerate gas cannot be achieved.

Another way to reduce the size and mass of a neutrino star is to consider new long-range interactions between neutrinos which are much stronger than gravity. In the case of Yukawa forces due to a new light scalar boson ϕ\phi with the coupling constant to neutrinos, yy, the Newton coupling GNG_{N} should be substituted in (1) by

Gν=y24π1mν2.G_{\nu}=\frac{y^{2}}{4\pi}\frac{1}{m_{\nu}^{2}}. (3)

As a result, we obtain

R=4π21ymν=17.8ymν.R=4\pi\sqrt{2}\frac{1}{ym_{\nu}}=\frac{17.8}{ym_{\nu}}. (4)

Notice that this expression does not depend on the mass of mediator mϕm_{\phi} as long as mϕ1/Rm_{\phi}\ll 1/R. For y=107y=10^{-7} and mν=0.05m_{\nu}=0.05 eV, one finds from (3) Gν=0.56\sqrt{G_{\nu}}=0.56 MeV-1 (that is, 22 orders of magnitude larger than GN\sqrt{G_{N}}), and correspondingly Eq. (4) gives R=0.7R=0.7 km.

Properties and formation of the ν\nu-clusters due to new scalar interactions were studied in Stephenson et al. (1998); Stephenson and Goldman (1993) using equations of Quantum Hadrodynamics. The central notion is the effective neutrino mass m~ν\tilde{m}_{\nu} in the background of degenerate neutrino gas and scalar field. The equation for m~ν\tilde{m}_{\nu} was obtained in the limit of static and uniform medium. This non-linear equation depends on the Fermi momentum kFk_{F} and the strength of interactions which is defined as

Gϕy2ω2π2mϕ2.G_{\phi}\equiv\frac{y^{2}\omega}{2\pi^{2}m_{\phi}^{2}}.

Here ω\omega is the number of degrees of freedom. For large densities before neutrino clustering, the effective mass is close to zero Stephenson et al. (1998):

m~ν2mνGϕkF2.\tilde{m}_{\nu}\approx\frac{2m_{\nu}}{G_{\phi}k_{F}^{2}}.

At small densities: m~νmν\tilde{m}_{\nu}\rightarrow m_{\nu}. Using m~ν\tilde{m}_{\nu}, the energy density of neutrinos, ρν\rho_{\nu}, and the energy density of the scalar field, ρϕ\rho_{\phi}, were computed in Stephenson et al. (1998) which gives the total energy per neutrino

ϵtot=ϵν+ϵϕ=1nF(ρν+ρϕ),\epsilon^{\rm tot}=\epsilon_{\nu}+\epsilon_{\phi}=\frac{1}{n_{F}}(\rho_{\nu}+\rho_{\phi}),

here nF=kF3/6π2n_{F}=k_{F}^{3}/6\pi^{2}. For large enough strength GϕG_{\phi} the energy ϵtot\epsilon^{\rm tot} as a function of kFk_{F} acquires a minimum with ϵtot<mν\epsilon^{\rm tot}<m_{\nu} around kFmin0.8mνk_{F}^{\min}\approx 0.8m_{\nu}. Clustering of neutrinos starts when kFk_{F} decreases due to the expansion of the Universe down to

kFkFmin.k_{F}\sim k_{F}^{\min}.

It is also assumed that with the expansion (decrease of nn), the kinetic energy could be converted into the increasing effective mass m~ν\tilde{m}_{\nu} and the gas became strongly degenerate.

To describe finite size objects, the static equations of motion were used with non-zero spatial derivatives of the fields, and consequently, m~ν\tilde{m}_{\nu}. It was assumed that the distribution of clusters on the number of neutrinos NN follows the density fluctuations and has the Harrison-Zeldovich spectrum. In Stephenson et al. (1998), the neutrino mass mν=13m_{\nu}=13 eV motivated by the Tritium experiment anomaly was used. Very small mediator masses mϕ1017m_{\phi}\sim 10^{-17} eV and couplings g1014g\sim 10^{-14} were taken. Consequently, the sizes of clusters equal 1013\sim 10^{13} cm, which is roughly the size of the Earth orbit, and the central densities could reach 101510^{15} cm-3.

Later the bound systems of heavy Dirac fermions (“nuggets”) due to Yukawa couplings with light or massless scalar were explored Wise and Zhang (2014, 2015); Gresham et al. (2017, 2018). With fermion masses mD100m_{D}\sim 100 GeV and coupling αϕ(0.010.1)\alpha_{\phi}\sim(0.01-0.1) such systems were applied to the Asymmetric Dark Matter (ADM). The description of nuggets in Wise and Zhang (2015), Gresham et al. (2017) is similar to that of neutrino stars Stephenson et al. (1998)111 The authors of those ADM papers overlooked (and do not refer to) similar papers on neutrinos Stephenson et al. (1998) published 20 years before. In the first version of our paper we, being focussed on neutrinos, overlooked the papers on ADM.. The effective mass of fermion was introduced as in Stephenson et al. (1998) and equations for its dependence on rr was derived. The system of equations for the scalar field and Fermi momentum of ADM particles pF(r)p_{F}(r) (and therefore their density) was presented. In Wise and Zhang (2015) the equations were solved by introducing an ansatz for pF(r)p_{F}(r), thus reducing the system to a single equation. In Gresham et al. (2017) the complete system of equations was solved numerically. Dependence of properties of nuggets on NN, in particular R(N)R(N), were studied. It was established in Wise and Zhang (2015) that with the increase of NN the radius first decreases, reaches minimum and then increases. The increase is associated to transition from non-relativistic to relativistic regime. While in Wise and Zhang (2015) the mediator was assumed to be massless, in Gresham et al. (2017) dependence of characteristics of nuggets on mϕm_{\phi} was studied and, in particular, the case of “saturation”, which corresponds to R>1/mϕR>1/m_{\phi}, was noticed.

Due to strongly different values of masses and coupling constant the formation and cosmological consequences of nuggets are completely different from those of neutrino stars. Also observational methods and perspectives of detection are different. Our paper is continuation of the work in Stephenson et al. (1998) for neutrinos. Where possible, we compare our results with those for ADM.

We perform a systematic study of various aspects of physics of neutrino clusters. The paper contains a number of derivations and auxiliary material which will be useful for further studies and applications.

In this paper, we revisit the possibility of formation of the neutrino bound states and bound systems due to neutrino interactions with very light scalar bosons. First the bound states of two neutrinos are considered. We formulate conditions for their existence and find their parameters such as eigenstates and eigenvalues. For usual neutrinos and allowed values of yy and mϕm_{\phi}, the sizes of these bound states are of the astronomical distances which has no sense. For sterile neutrinos with mass mν10m_{\nu}\geq 10 keV and y>104y>10^{-4} the size can be in sub-cm range and formation of such states may have implications for dark matter.

Then we study properties and formation of the NN-neutrino bound systems — the ν\nu-clusters with a sufficiently large NN, so that the system can be quantitatively described using Fermi-Dirac statistics. We compute characteristics of possible stable final configurations of the ν\nu-clusters as function of NN — mostly in the state of degenerate Fermi gas (ground state) but also with thermal distributions. Energy loss of the ν\nu-clusters is estimated. Depending on NN (we explore the entire possible range) the clusters can be in non-relativistic or relativistic regimes and we study the transition between these two regimes. In the non-relativistic case, using equations of motions for ν\nu and ϕ\phi, we rederive the Lane-Emden equation for the neutrino density. We obtain qualitative analytic results for parameters of the neutrino cluster. The equation for the relativistic case is derived which is similar to the one used in Stephenson et al. (1998). We use presently known values of neutrino masses which are two orders of magnitude smaller than in Stephenson et al. (1998), and consequently, parameters of the clouds are different.

Understanding the formation of neutrino clusters from homogeneous relativistic neutrino gas requires numerical simulations. In this connection we present some relevant analytic results. One possible mechanism of formation is the development of instabilities in the uniform neutrino background, which leads to fragmentation and further redistribution of neutrino density, thus approaching final degenerate configurations. In some parameter space, one can use an analogy between Yukawa forces and gravity and therefore compare it with the formation of dark matter halos due to gravity.

The paper is organized as follows. We start with re-derivation of the Yukawa potential in Sec. II, where we adopt the approach developed in Ref. Smirnov and Xu (2019) and generalize it to the relativistic case. In Sec. III, we study the two-body bound states in the framework of quantum mechanics. In Sec. IV, the NN-body bound systems are considered. The results of Sec. III and Sec. IV apply to the non-relativistic neutrinos, while a study of the relativistic case is presented in Sec. V. In Sec. VI, we speculate on possible mechanisms of formation of the ν\nu-clusters which include fragmentation of the cosmic neutrino background induced by the expansion of the Universe. Further evolution and formation of final configurations are outlined. Energy loss of the clusters is estimated. Analogy with formation of DM halos is outlined. Finally, we conclude in Sec. VII. Various details and explanations are presented in the appendices.

II Yukawa potential for neutrinos

In general, to form a bound state, the range of the force, RforceR_{{\rm force}}, needs to be longer than the de Broglie wavelength of the neutrino, λν\lambda_{\nu}:

Rforce>λν.R_{{\rm force}}>\lambda_{\nu}. (5)

For a non-relativistic neutrino λν=(mνv)1\lambda_{\nu}=(m_{\nu}v)^{-1}, where vv is the neutrino velocity, and Rforce1/nϕR_{{\rm force}}\sim 1/n_{\phi} is given by the inverse of the mediator mass. Therefore Eq. (5) can be written as mϕ<mνvm_{\phi}<m_{\nu}v, which implies that mediator should be lighter than the neutrino to form a bound state222In the standard model (SM), the ZZ boson does mediate an attractive force between a neutrino and an antineutrino but due to its large mass, the range of this force is too short to bind neutrinos together..

Let us consider a light real scalar boson ϕ\phi interacting with neutrinos

=12μϕμϕ12mϕ2ϕ2+ν¯i∂̸νmνν¯νyν¯ϕν.{\cal L}=\frac{1}{2}\partial^{\mu}\phi\partial_{\mu}\phi-\frac{1}{2}m_{\phi}^{2}\phi^{2}+\overline{\nu}i\not{\partial}\nu-m_{\nu}\overline{\nu}\nu-y\overline{\nu}\phi\nu. (6)

We assume mϕmνm_{\phi}\ll m_{\nu}, so that ν\nu can be treated as point-like particles while ϕ\phi as a classical field. This is valid as long as we are not concerned with hard scattering processes. In this regime, the effect of the ϕ\phi field on ν\nu particles can be described using Yukawa potential.

From Eq. (6), we obtain the Euler-Lagrange equations of motion (EOM) for ν\nu and ϕ\phi:

i∂̸ν(mν+yϕ)ν\displaystyle i\not{\partial}\nu-(m_{\nu}+y\phi)\nu =0,\displaystyle=0\thinspace, (7)
(2+mϕ2)ϕ+yν¯ν\displaystyle(\partial^{2}+m_{\phi}^{2})\phi+y\overline{\nu}\nu =0.\displaystyle=0\thinspace. (8)

According to Eq. (7) the effect of the ϕ\phi field on the motion of ν\nu can be accounted by shifting its mass:

m~νmν+V,Vyϕ.\tilde{m}_{\nu}\equiv m_{\nu}+V,\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ V\equiv y\phi\thinspace. (9)

Here VV is the potential, and m~ν\tilde{m}_{\nu} is treated as the effective mass of neutrino in background.

At quantum level ν¯ν\overline{\nu}\nu is replaced by the expectation value computed as (see Appendix A):

ν¯νν¯ν=d3𝐩(2π)3m~νE𝐩f(t,𝐱,𝐩),\overline{\nu}\nu\rightarrow\langle\overline{\nu}\nu\rangle=\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{\tilde{m}_{\nu}}{E_{\mathbf{p}}}f(t,\ \mathbf{x},\ \mathbf{p})\thinspace, (10)

where E𝐩=m~ν2+|𝐩|2E_{\mathbf{p}}=\sqrt{\tilde{m}_{\nu}^{2}+|\mathbf{p}|^{2}} and f(t,𝐱,𝐩)f(t,\ \mathbf{x},\ \mathbf{p}) is the neutrino distribution function normalized in such a way that the number density, nn, and the total number of ν\nu, NN, are given by

n=f(t,𝐱,𝐩)d3𝐩(2π)3,N=n(t,𝐱)d3𝐱.n=\int f(t,\ \mathbf{x},\ \mathbf{p})\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\thinspace,\ \ N=\int n(t,\ \mathbf{x})d^{3}\mathbf{x}\thinspace.

In the non-relativistic case, m~ν/E𝐩1\tilde{m}_{\nu}/E_{\mathbf{p}}\approx 1, Eq. (10) gives

ν¯νn.\langle\overline{\nu}\nu\rangle\approx n. (11)

We assume a static distribution, so that tϕ=0\partial_{t}\phi=0. Replacing ψ¯ψ\overline{\psi}\psi by ψ¯ψn\langle\overline{\psi}\psi\rangle\approx n in Eq. (8) and using relation ϕ=V/y\phi=V/y, (see Eq. (9)) we obtain equation for the potential

[2mϕ2]V=y2n.\left[\nabla^{2}-m_{\phi}^{2}\right]V=y^{2}n. (12)

For spherically symmetric distribution of neutrinos, n=n(r)n=n(r), where rr is the distance from the center, Eq. (12) can be solved analytically Smirnov and Xu (2019), giving

V=y2mϕr[emϕr0r𝑑rrn(r)sinh(mϕr)+sinh(mϕr)r𝑑remϕrrn(r)].V=-\frac{y^{2}}{m_{\phi}r}\left[e^{-m_{\phi}r}\int_{0}^{r}dr^{\prime}r^{\prime}n(r^{\prime})\sinh(m_{\phi}r^{\prime})+\sinh(m_{\phi}r^{\prime})\int_{r}^{\infty}dr^{\prime}e^{-m_{\phi}r^{\prime}}r^{\prime}n(r^{\prime})\right]. (13)

For n=Nδ3(𝐱)n=N\delta^{3}(\mathbf{x}) (all neutrinos are in origin) the solution reproduces the Yukawa potential

V(r)=Ny24π1rermϕ.V(r)=-N\frac{y^{2}}{4\pi}\frac{1}{r}e^{-rm_{\phi}}. (14)

Note that VV is the potential experienced by a given neutrino, and for two neutrinos separated by a distance rr, the total binding energy is given by Eq. (14) with N=1N=1 (see Appendix B for details).

For relativistic neutrinos, if tϕ\partial_{t}\phi can be neglected333The time derivative tϕ\partial_{t}\phi can only be neglected for stationary (e.g. the ground state of a binary system) or approximately stationary distributions (e.g., a many-body system with a low temperature). Otherwise, tϕ\partial_{t}\phi plays an important role in quantum state transitions or in energy loss of a many-body system due to emission of ϕ\phi waves. This is similar to the familiar example of an excited atom that will undergo transition to the ground state in finite time via spontaneous emission of electromagnetic waves., one should use general expression in (10) and the equation for the potential becomes

[2mϕ2]V=y2d3𝐩(2π)3m~νE𝐩f(𝐱,𝐩).\left[\nabla^{2}-m_{\phi}^{2}\right]V=y^{2}\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{\tilde{m}_{\nu}}{E_{\mathbf{p}}}f(\mathbf{x},\ \mathbf{p}). (15)

The resulting VV is suppressed by a factor of m~ν/E𝐩\tilde{m}_{\nu}/E_{\mathbf{p}} in comparison to the non-relativistic case. This equation can be written in the form Eq. (12) with nn substituted by the effective number density, n~\tilde{n}, defined as

n~d3𝐩(2π)3m~νE𝐩f(𝐱,𝐩).\tilde{n}\equiv\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{\tilde{m}_{\nu}}{E_{\mathbf{p}}}f(\mathbf{x},\ \mathbf{p}).

In the non-relativistic case n~n\tilde{n}\rightarrow n.

In this paper we focus on the real scalar field. Pseudo-scalar mediators (e.g. the majoron) cause spin-dependent forces, and the corresponding potential can be found in Chikashige et al. (1981). For a two-body system, the potential might be able to bind the pair if the mediator mass is sufficiently light. In an NN-body system with large NN, the effect of spin-dependent forces is suppressed by spin averaging.

In the case of a complex scalar, our analysis can be applied to its real part, while the effect of its imaginary part, which is a pseudo-scalar, can be neglected if the two parts have the same mass. In models with spontaneous symmetry breaking such as the one in Chikashige et al. (1981), only the pseudo-scalar component is massless and therefore leads to long-range effects.

III Two-neutrino bound states

In this section, we consider bound states of two non-relativistic neutrinos due to the attractive ϕ\phi potential. For DM particles this was considered in a number of papers before (see Wise and Zhang (2014) and references therein).

The strength of the interaction required to form bound states can be estimated in the following way. If the wave function of the bound state is localized in the region of radius RR, then according to the Heisenberg uncertainty principle the momentum of neutrino is p1/Rp\sim 1/R, and consequently, the kinetic energy, Ekinp2/2mνE_{kin}\approx p^{2}/2m_{\nu}, equals

Ekin12mν1R2.E_{{\rm kin}}\sim\frac{1}{2m_{\nu}}\frac{1}{R^{2}}. (16)

The condition of neutrino trapping in the potential VV reads

EkinV,(r<R).E_{{\rm kin}}\lesssim-V,\ \,\,(r<R). (17)

Plugging EkinE_{{\rm kin}} from Eq. (16) and VV from Eq. (14) with N=1N=1 into this equation we obtain

12mνy2R4πeRmϕ.\frac{1}{2m_{\nu}}\lesssim\frac{y^{2}R}{4\pi}e^{-Rm_{\phi}}. (18)

Maximal value of the r.h.s. of this equation equals y2/(4πemϕ)y^{2}/(4\pi em_{\phi}), which is achieved at R=mϕ1R=m_{\phi}^{-1}. Then, according to (18), the sufficient condition for existence of bound state can be written as

λy28πmνmϕ0.7,\lambda\equiv\frac{y^{2}}{8\pi}\frac{m_{\nu}}{m_{\phi}}\gtrsim 0.7, (19)

where λ\lambda can be interpreted as strength of interaction. Quantitative study below gives λ>0.84\lambda>0.84. The condition can be rewritten as the upper bound on mϕm_{\phi}:

mϕy24πmν=41017(y107)2eV.m_{\phi}\lesssim\frac{y^{2}}{4\pi}m_{\nu}=4\cdot 10^{-17}\left(\frac{y}{10^{-7}}\right)^{2}{\rm eV}.

III.1 Quantitative study

Refer to caption
Figure 1: The radial dependences of the wave functions of neutrino bound states for different values of the interaction strength λ\lambda defined in Eq. (19). In all the panels we take the orbital momentum l=0l=0. For λ=0.5\lambda=0.5, since no bound states can be formed, we plot the wave function with the lowest energy.

Quantitative treatment of neutrino bound states requires solution of the two-body Schrödinger equation. Let 𝐫1\mathbf{r}_{1} and 𝐫2\mathbf{r}_{2} be the coordinates of two neutrinos. The potential depends on the relative distance between neutrinos 𝐫𝐫1𝐫2\mathbf{r}\equiv\mathbf{r}_{1}-\mathbf{r}_{2} only. Consequently, the wave function of bound state depends on 𝐫\mathbf{r}, ν=ν(𝐫)\nu=\nu(\mathbf{r}), and it obeys the Schrödinger equation

[1mν2+V(r)]ν(𝐫)=Eν(𝐫),\left[-\frac{1}{m_{\nu}}\nabla^{2}+V(r)\right]\nu(\mathbf{r})=E\nu(\mathbf{r}),\\ (20)

with V(r)=V(r,N=1)V(r)=V(r,N=1) given in Eq. (14), r=|𝐫|r=|\mathbf{r}| and EE being the energy eigenvalue. The numeric coefficient of the kinetic term takes into account the reduced mass of neutrino (see Appendix C). Since the potential V(r)V(r) is independent of the direction, the wave function can be factorized as

ν=u(r)rYlm(θ,φ).\nu=\frac{u(r)}{r}Y_{l}^{m}(\theta,\ \varphi). (21)

Here u(r)u(r) is a radial component and Ylm(θ,φ)Y_{l}^{m}(\theta,\ \varphi) is a spherical harmonic function. Plugging Eq. (21) into Eq. (20), we obtain the equation for u(r)u(r):

1mνd2u(r)dr2+[V(r)+1mνl(l+1)r2]u(r)=Eu(r).-\frac{1}{m_{\nu}}\frac{d^{2}u(r)}{dr^{2}}+\left[V(r)+\frac{1}{m_{\nu}}\frac{l(l+1)}{r^{2}}\right]u(r)=Eu(r).

For zero angular momentum, l=0l=0, the equation simplifies further:

u′′(r)+mνV(r)u(r)=mνEu(r).-u^{\prime\prime}(r)+m_{\nu}V(r)u(r)=m_{\nu}Eu(r). (22)

We solve this equation numerically (see details in Appendix C). u(r)u(r) for different values of λ\lambda are shown in Fig. 1. According to this figure there is no bound state for λ=0.5\lambda=0.5: the solution with the lowest energy level does not converge to zero when rr\rightarrow\infty. For other specific values, λ=0.9\lambda=0.9, 3.63.6, and 9.09.0, the figure shows one, two, and three bound states. With further increase of λ\lambda, the number of allowed energy levels of increases. Varying λ\lambda we find that for

  • λ<0.84\lambda<0.84, no bound states can be formed;

  • 0.84<λ<3.20.84<\lambda<3.2, there is one bound state (the 1s state);

  • 3.2<λ<7.23.2<\lambda<7.2, two bound states with l=0l=0 (the 2s state) can be formed;

  • λ>7.2\lambda>7.2, three or more bound states exist, including states with l0l\neq 0.

III.2 Coulomb limit

The limit of mϕ0m_{\phi}\rightarrow 0 (λ\lambda\rightarrow\infty) corresponds to a Coulomb-like potential, and in this case, the Schrödinger equation (22) can be solved analytically. There is an infinite number of levels and the wave function of ground state equals

u(r)=2R03/2rexp(rR0),u(r)=2R_{0}^{3/2}r\exp\left(-\frac{r}{R_{0}}\right), (23)

where

R08πmνy2R_{0}\equiv\frac{8\pi}{m_{\nu}y^{2}}\thinspace (24)

is the radius of the bound state, and the corresponding eigenvalue equals

E=y4mν64π2.E=-\frac{y^{4}m_{\nu}}{64\pi^{2}}\thinspace. (25)

This EE can be interpreted as the binding energy of the state. For non-zero mϕm_{\phi} such that R0mϕ1R_{0}\ll m_{\phi}^{-1} (equivalent to λ1\lambda\gg 1), the range of the force is much larger than the radius of the bound state in the Coulomb limit. In this case, Eq. (24) and (25) can be used as a good approximation. The correction to the Coulomb limit result due to non-zero mϕm_{\phi} can be computed using the variation principle (see below).

III.3 Approximate solutions for non-zero mass of scalar from the variation principle

According to the variation principle the expectation of the Hamiltonian

Hu(r)Hu(r)𝑑ru(r)u(r)𝑑r\langle H\rangle\equiv\frac{\int u^{*}(r)Hu(r)\thinspace dr}{\int u^{*}(r)u(r)\thinspace dr}

reaches the minimal value when uu is the wave function of the ground state, while H\langle H\rangle itself gives the corresponding eigenvalue. The Hamiltonian for the radial wave function and l=0l=0 reads

H=1mνd2dr2+V(r).H=-\frac{1}{m_{\nu}}\frac{d^{2}}{dr^{2}}+V(r)\thinspace.

Assuming solution in the form Eq. (23): u(r)=2a3/2rer/au(r)=2a^{3/2}re^{-r/a} with aa being a free parameter, we find

H=1a2mν[1ay2mνπ(amϕ+2)2].\langle H\rangle=\frac{1}{a^{2}m_{\nu}}\left[1-\frac{ay^{2}m_{\nu}}{\pi(am_{\phi}+2)^{2}}\right]. (26)

Minimization of H\langle H\rangle with respect to aa, that is dH/da=0d\langle H\rangle/da=0, gives

(16πλ+amνy2)332πλ2amνy2(16πλ+3amνy2)=0,(16\pi\lambda+am_{\nu}y^{2})^{3}-32\pi\lambda^{2}am_{\nu}y^{2}(16\pi\lambda+3am_{\nu}y^{2})=0, (27)

which is a cubic equation with respect to aa. The cubic equation has one negative solution which can be ignored, and two others which can be either both complex (in this case the two solutions are conjugate to each other) or both real and positive (in this case the two solutions are identical). Only the later case leads to a physical solution and requires that the discriminant is positive:

λ2+209λ30.\lambda^{2}+\frac{20}{9}\lambda-3\geq 0\thinspace.

From this inequality we obtain

λ19(7710)0.95,\lambda\geq\frac{1}{9}\left(7\sqrt{7}-10\right)\approx 0.95\thinspace,

which is about 13% larger than the exact value 0.840.84 obtained in Sec. III.1. In Wise and Zhang (2014) the bound λ0.84\lambda\geq 0.84 was also obtained.

Using the variation principle in this way may not give an accurate wave function, but usually leads to a very accurate result for eigenvalues Griffiths (2016). Note that here we do not require that λ1mϕ\lambda^{-1}\propto m_{\phi} is perturbatively small. For very small λ1\lambda^{-1} the result is expected to be more accurate. The series expansion of the positive physical solution of Eq. (27) in λ1\lambda^{-1} gives

x=8π[1+34λ2+𝒪(1λ3)].x=8\pi\left[1+\frac{3}{4\lambda^{2}}+{\cal O}\left(\frac{1}{\lambda^{3}}\right)\right].

Therefore, the radius of the system R=a=x/y2mνR=a=x/y^{2}m_{\nu} equals

R8πmνy2[1+34λ2].R\approx\frac{8\pi}{m_{\nu}y^{2}}\left[1+\frac{3}{4\lambda^{2}}\right].

The eigenvalue is corrected according to (26) as

E=Hy4mν64π2[12λ+32λ2].E=\langle H\rangle\approx-\frac{y^{4}m_{\nu}}{64\pi^{2}}\left[1-\frac{2}{\lambda}+\frac{3}{2\lambda^{2}}\right]. (28)

The lowest order expressions for RR and binding energy (28) coincide with those in Wise and Zhang (2014).

Numerically, the radius equals

R=51011cm(0.1eVmν)(107y)2.R=5\cdot 10^{11}{\rm cm}\left(\frac{0.1{\rm eV}}{m_{\nu}}\right)\left(\frac{10^{-7}}{y}\right)^{2}. (29)

This is much larger than the distance between relic neutrinos (without clustering) 0.16\sim 0.16 cm, and therefore existence of bound states of such size has no practical sense.

For two-body bound states, it is impossible to enter the relativistic regime. Indeed, according to Eq. (25) or Eq. (28) the kinetic energy being of the order of binding energy equals y4mν/64π2y^{4}m_{\nu}/64\pi^{2}. Therefore the relativistic case, Ekin/mν1E_{\rm kin}/m_{\nu}\geq 1 implies y(64π2)1/4y\gtrsim(64\pi^{2})^{1/4} which would not only violate the perturbativity bound but also be excluded by various experimental limits — see Appendix D.

The results of this section may have applications for sterile neutrinos with mass mν10m_{\nu}\geq 10 keV. These neutrinos could compose dark matter of the Universe. For mν10m_{\nu}\geq 10 keV and y=103y=10^{-3} we find from (29) R=0.05R=0.05 cm.

IV NN-neutrino bound systems

Let us consider NN neutrinos trapped in the Yukawa potential with NN being large enough, so that neutrinos form a statistical system which is described by the Fermi-Dirac distribution function

f=1expEμT+1.f=\frac{1}{\exp\frac{E-\mu}{T}+1}. (30)

Here EE is the total neutrino energy, μ\mu is the chemical potential and TT is the temperature. Recall that for this distribution the number density, nn, the energy density, ρ\rho, and pressure 𝒫{\cal P} of the system equal respectively

n=fd3𝐩(2π)3,ρ=Efd3𝐩(2π)3,𝒫=13|𝐩|2Efd3𝐩(2π)3.n=\int f\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\thinspace,\,\,\,\,\rho=\int Ef\thinspace\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\thinspace,\,\,\,\,{\cal P}=\frac{1}{3}\int\frac{|\mathbf{p}|^{2}}{E}f\thinspace\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\thinspace.

The ground state of this system (T0T\rightarrow 0) is the degenerate Fermi gas with distribution f(p)=1f(p)=1, if E<μE<\mu, and f(p)=0f(p)=0 if E>μE>\mu according to Eq. (30). (The case of finite TT will be discussed in Sec. VI). Consequently, in the degenerate state

n=0pF4πp2dp(2π)3=pF36π2,n=\int_{0}^{p_{F}}\frac{4\pi p^{2}dp}{(2\pi)^{3}}=\frac{p_{F}^{3}}{6\pi^{2}}\thinspace, (31)

and the total number of neutrinos NN for uniform distribution in the sphere of radius RR equals

N=2pF39πR3.N=\frac{2p_{F}^{3}}{9\pi}R^{3}\thinspace. (32)

Eq. (31) applies to both relativistic and non-relativistic cases. For ρ\rho and 𝒫{\cal P}, the non-relativistic results are

𝒫deg=16π20pFp2Ep2𝑑p=pF530π2mν,ρ=32𝒫,{\cal P}_{\rm deg}=\frac{1}{6\pi^{2}}\int_{0}^{p_{F}}\frac{p^{2}}{E}p^{2}dp=\frac{p_{F}^{5}}{30\pi^{2}m_{\nu}}\thinspace,\,\,\,\ \rho=\frac{3}{2}{\cal P}, (33)

while the relativistic results can be obtained by computing the integrals numerically.

IV.1 Qualitative estimation

Let us consider NN neutrinos uniformly distributed in a sphere of radius RR. The Yukawa attraction produces a pressure, 𝒫Yuk{\cal P}_{{\rm Yuk}}, which for mϕ=0m_{\phi}=0 can be estimated as

𝒫Yuk1𝒱[12y2N24πR],{\cal P}_{{\rm Yuk}}\simeq\frac{1}{{\cal V}}\left[-\frac{1}{2}\frac{y^{2}N^{2}}{4\pi R}\right]\thinspace, (34)

where 𝒱=43πR3{\cal V}=\frac{4}{3}\pi R^{3} is the volume and the quantity in the brackets is the total potential energy.

Static configuration corresponds to the equilibrium between this Yukawa pressure and the pressure of degenerate fermion gas (Pauli repulsion):

𝒫deg𝒫Yuk.{\cal P}_{{\rm deg}}\simeq{\cal P}_{{\rm Yuk}}\thinspace.

Using Eq. (33) and Eq. (34) this equality can be rewritten as

pF530π2mν3y2N22(4π)2R4.\frac{p_{F}^{5}}{30\pi^{2}m_{\nu}}\simeq\frac{3y^{2}N^{2}}{2\left(4\pi\right)^{2}R^{4}}\thinspace. (35)

Inserting Eq. (32) in Eq. (35) we find

R8ymνpF.R\simeq\frac{8}{y\sqrt{m_{\nu}p_{F}}}\thinspace. (36)

Alternatively, we can express pFp_{F} in terms of NN and obtain

R30mνy2(1N)1/3.R\simeq\frac{30}{m_{\nu}y^{2}}\left(\frac{1}{N}\right)^{1/3}\thinspace. (37)

Plugging RR from Eq. (37) back into Eq. (32), we obtain a relation between NN and pFp_{F}:

N40y3(pFmν)3/2.N\simeq\frac{40}{y^{3}}\left(\frac{p_{F}}{m_{\nu}}\right)^{3/2}\thinspace. (38)

Notice that for N2N\rightarrow 2, Eq. (37) recovers approximately the result in Eq. (24) for the two-body system. A noteworthy feature of Eq. (37) is that the radius decreases with NN: RN1/3R\propto N^{-1/3}. This implies that increasing NN leads to a more compact cluster. This dependence is valid only in the non-relativistic regime. When NN is too large, the Fermi momentum pFp_{F} which increases with NN according to Eq. (38) will exceed mνm_{\nu} and the system enters the relativistic regime. As we will show in Sec. V, in the relativistic regime, the attractive force becomes suppressed, causing RR to increases with NN.

IV.2 Quantitative study

In an object of finite size, the density and pressure depend on distance from the center of object. The dependence can be obtained using the differential (local) form of the equality of forces, which is given by the equation of hydrostatic equilibrium. For degenerate neutrino gas it is reduced to the Lane-Emden (L-E) equation Chandrasekhar and Chandrasekhar (1957). The L-E equation can be also derived from equations of motion of the fields (see Appendix F).

The equation of hydrostatic equilibrium expresses the equality of the repulsive force due to the degenerate pressure, FdegF_{{\rm deg}}, and the Yukawa attractive force, FYukF_{\text{Yuk}}, acting on a unit of volume at distance rr from the center:

Fdeg(r)d𝒫deg(r)dr=FYuk(r).F_{{\rm deg}}(r)\equiv\frac{d{\cal P}_{{\rm deg}}(r)}{dr}=F_{\text{Yuk}}(r)\thinspace. (39)

For non-relativistic neutrinos, 𝒫deg{\cal P}_{{\rm deg}} is given in Eq. (33), and using Eq. (31), 𝒫deg{\cal P}_{{\rm deg}} can be rewritten as

𝒫deg=Kn5/3,K(6π2)5/330π2mν=(6π2)2/35mν.{\cal P}_{{\rm deg}}=Kn^{5/3},\,\,\,\ K\equiv\frac{(6\pi^{2})^{5/3}}{30\pi^{2}m_{\nu}}=\frac{(6\pi^{2})^{2/3}}{5m_{\nu}}\thinspace. (40)

The Yukawa force equals

FYuk(r)=y24πr2n(r)N(r),F_{{\rm Yuk}}(r)=-\frac{y^{2}}{4\pi r^{2}}n(r)N(r)\thinspace, (41)

where

N(r)=0r𝑑r4πrn2(r)N(r)=\int_{0}^{r}dr^{\prime}4\pi r^{\prime}{}^{2}n(r^{\prime})

is the number of neutrinos inside a sphere of radius rr. Inserting Eqs. (40) and (41) into Eq. (39) gives the equation for n(r)n(r):

53Kn2/3dndr=y2r2n0r𝑑rr2n(r).\frac{5}{3}Kn^{2/3}\frac{dn}{dr}=\frac{y^{2}}{r^{2}}n\int_{0}^{r}dr^{\prime}r^{\prime 2}n(r^{\prime})\thinspace.

Dividing this equation by n/r2n/r^{2} and then differentiating by rr, we obtain

52Kddr[r2dn2/3dr]=y2r2n,\frac{5}{2}K\frac{d}{dr}\left[r^{2}\frac{dn^{2/3}}{dr}\right]=-y^{2}r^{2}n\thinspace,

or

1r2ddr[r2dn2/3dr]=y2κn,\frac{1}{r^{2}}\frac{d}{dr}\left[r^{2}\frac{dn^{2/3}}{dr}\right]=-y^{2}\kappa n\thinspace, (42)

where

κ2mν(6π2)2/3.\kappa\equiv\frac{2m_{\nu}}{(6\pi^{2})^{2/3}}\thinspace.

Eq. (42) is nothing but the Lane-Emden equation Chandrasekhar and Chandrasekhar (1957) for specific values of parameters. In particular, γ\gamma, defined as the ratio of powers of nn in the r.h.s. and the l.h.s. of Eq. (42), equals γ=1/(2/3)=3/2\gamma=1/(2/3)=3/2. For this value the NN-body system has a finite size RR defined by

n(R)=0.n(R)=0.

We solve the L-E equation (42) numerically with the boundary condition

n(r=0)=n0,orequivalently,pF(r=0)=pF0.n(r=0)=n_{0},\leavevmode\nobreak\ \leavevmode\nobreak\ {\rm or\leavevmode\nobreak\ \leavevmode\nobreak\ equivalently},\leavevmode\nobreak\ \leavevmode\nobreak\ p_{F}(r=0)=p_{F0}.

Fig. 2 shows solutions of n(r)n(r) for different values of pF0/mνp_{F0}/m_{\nu}. As rr increases, the number densities drop down to zero at

R19.9ymνpF0.R\approx\frac{19.9}{y\sqrt{m_{\nu}p_{F0}}}. (43)

Integrating n(r)n(r) with rr ranging from 0 to RR gives the total number of neutrinos:

N93.6y3(pF0mν)3/2.N\approx\frac{93.6}{y^{3}}\left(\frac{p_{F0}}{m_{\nu}}\right)^{3/2}\thinspace. (44)

Combining Eqs. (43) and (44), we obtain expression for the radius

R90.4mνy2(1N)1/3,R\approx\frac{90.4}{m_{\nu}y^{2}}\left(\frac{1}{N}\right)^{1/3}\thinspace, (45)

which shows that RR decreases with NN, in agreement with the qualitative estimations in Sec. IV.1. The results in Eqs. (43), (44) and (45) have the same parametric dependence as Eqs. (36), (38) and (37) and differ by numerical factors 2.5, 2.3 and 3 correspondingly.

Notice that for the Fermi momentum pF0mνp_{F0}\sim m_{\nu} the Eq. (43) gives R=19.9/ymνR=19.9/ym_{\nu}, which practically coincides with (4) obtained from rescaling of parameters of neutron star. The coincidence is not accidental since parameters of neutron star correspond to pFmnp_{F}\sim m_{n}. Other characteristics of a ν\nu-cluster for pF0mνp_{F0}\sim m_{\nu} are Mν=31011M_{\nu}=3\cdot 10^{-11} g (for y=107y=10^{-7}) and n0=2.6108n_{0}=2.6\cdot 10^{8} cm-3. The central density is determined according to Eq. (31) as

n0mν3=16π2(pF0mν)3.\frac{n_{0}}{m_{\nu}^{3}}=\frac{1}{6\pi^{2}}\left(\frac{p_{F0}}{m_{\nu}}\right)^{3}. (46)
Refer to caption
Figure 2: Dependence of neutrino number density n(r)n(r) on distance from center for different values of central density pF0/mνp_{F0}/m_{\nu}. We take y=107y=10^{-7} and mν=0.1m_{\nu}=0.1 eV.

In Eq. (42) both y2y^{2} and κ\kappa can be absorbed by the redefinition of rr:

r=ryκrymν.r^{\prime}=ry\sqrt{\kappa}\propto ry\sqrt{m_{\nu}}.

In terms of rr^{\prime}, Eq. (42) can be written as the universal form

1r2ddr[r2dn2/3dr]=n,\frac{1}{r^{\prime 2}}\frac{d}{dr^{\prime}}\left[r^{\prime 2}\frac{dn^{2/3}}{dr^{\prime}}\right]=-n\thinspace,

and its solution depends on the boundary condition n(0)n(0) which is free parameter restricted from above by pF0<mνp_{F0}<m_{\nu}. Thus, solution for different values of yy and mνm_{\nu} can be obtained from rescaling of rr:

r=r7(107y)(0.1eVmν)1/2,r=r_{7}\left(\frac{10^{-7}}{y}\right)\left(\frac{0.1{\rm eV}}{m_{\nu}}\right)^{1/2}, (47)

where r7r(y=107)r_{7}\equiv r(y=10^{-7}). The radius of a cluster is rescaled as in (47). For y=1014y=10^{-14} this equation gives R1012R\sim 10^{12} cm, i.e. orders of magnitude smaller than the Earth orbit. For the ν\nu-cluster made of ν2\nu_{2} (in the case of mass hierarchy) the radius is 6 times larger. The mass of such a cluster equals 310103\cdot 10^{10} g. The central density, and consequently, the mass of ν\nu-cluster is an independent parameter restricted by the neutrino mass. This consideration matches results in Eqs. (43) and (44).

IV.3 Non-degenerate ν\nu-clusters

As we will show in Sec. I the energy loss (cooling) of ν\nu-clusters due to ϕ\phi bremsstrahlung is negligible. Therefore the final configuration of degenerate neutrino gas may not be achieved. In this connection we consider the non-degenerate neutrino system characterized by a non-zero temperature TT. The number density equals

nT=12π20p2dpep/T+1=I32π2T3,n_{T}=\frac{1}{2\pi^{2}}\int_{0}^{\infty}\frac{p^{2}dp}{e^{p/T}+1}=\frac{I_{3}}{2\pi^{2}}T^{3},

where

In0dxxn1ex+1,I_{n}\equiv\int_{0}^{\infty}\frac{dxx^{n-1}}{e^{x}+1}, (48)

and I3I_{3} is related to the Riemann zeta function: I3=3ζ(3)/2=1.803I_{3}=3\zeta(3)/2=1.803.

For simplicity we assume that neutrinos are distributed uniformly in the sphere of radius RTR_{T}. Then

RT=(3N4πnT)1/3=(3π2I3)1/3N1/3T.R_{T}=\left(\frac{3N}{4\pi n_{T}}\right)^{1/3}=\left(\frac{3\pi}{2I_{3}}\right)^{1/3}\frac{N^{1/3}}{T}. (49)

If neutrinos are non-relativistic, the total kinetic energy can be computed as

EK4π3RT312π20p4dp2mν1ep/T+1=I52I3T2mνN,E_{K}\approx\frac{4\pi}{3}R_{T}^{3}\frac{1}{2\pi^{2}}\int^{\infty}_{0}\frac{p^{4}dp}{2m_{\nu}}\frac{1}{e^{p/T}+1}=\frac{I_{5}}{2I_{3}}\frac{T^{2}}{m_{\nu}}N, (50)

where I5=45ζ(5)/2=23.33I_{5}=45\zeta(5)/2=23.33 [see (48)] and we used Eq. (49) for RTR_{T}.

The potential energy of a cluster equals

EV=y24π0RT(4πnTr3/3)(4πnTr2dr)4πr=415πnT2RT5y2=y2N28πRT,E_{V}=-\frac{y^{2}}{4\pi}\int_{0}^{R_{T}}\frac{(4\pi n_{T}r^{3}/3)(4\pi n_{T}r^{2}dr)}{4\pi r}=-\frac{4}{15}\pi n_{T}^{2}R_{T}^{5}y^{2}=-\frac{y^{2}N^{2}}{8\pi R_{T}}, (51)

or eliminating RTR_{T}:

EV=320π(2I33π)1/3y2TN5/3.E_{V}=-\frac{3}{20\pi}\left(\frac{2I_{3}}{3\pi}\right)^{1/3}y^{2}TN^{5/3}. (52)

The ratio of the kinetic and potential energies can be written as

ξEEK|EV|=10I53(I3/π)4/3(32)1/3Ty2mνN2/3.\xi_{E}\equiv\frac{E_{K}}{|E_{V}|}=\frac{10I_{5}}{3(I_{3}/\pi)^{4/3}}\left(\frac{3}{2}\right)^{1/3}\leavevmode\nobreak\ \frac{T}{y^{2}m_{\nu}N^{2/3}}. (53)

The ratio ξE\xi_{E} in Eq. (53) depends on the temperature TT. Therefore, given a certain value of ξE\xi_{E}, it allows to determine TT:

T=15(32)2/3(I3π)4/3I51y2N2/3ξEmν.T=\frac{1}{5}\left(\frac{3}{2}\right)^{2/3}\left(\frac{I_{3}}{\pi}\right)^{4/3}I_{5}^{-1}y^{2}N^{2/3}\xi_{E}m_{\nu}.

For TpF/3T\rightarrow p_{F}/3 it matches Eq. (44), numerically:

T=0.081mν(y107)2(N61022)2/3ξE=0.081mν(y3N60)2/3ξE.T=0.081m_{\nu}\left(\frac{y}{10^{-7}}\right)^{2}\left(\frac{N}{6\cdot 10^{22}}\right)^{2/3}\xi_{E}=0.081m_{\nu}\left(\frac{y^{3}N}{60}\right)^{2/3}\xi_{E}. (54)

Then, according to (49) the radius equals

RT=1mν(23)1/35I5(I3/π)5/31y2N1/3ξE,R_{T}=\frac{1}{m_{\nu}}\left(\frac{2}{3}\right)^{1/3}\frac{5I_{5}}{(I_{3}/\pi)^{5/3}}\leavevmode\nobreak\ \frac{1}{y^{2}N^{1/3}\xi_{E}},

and numerically

RT=1.33km(107y)2(61022N)1/3(1ξE)(0.1eVmν).R_{T}=1.33\leavevmode\nobreak\ {\rm km}\left(\frac{10^{-7}}{y}\right)^{2}\left(\frac{6\cdot 10^{22}}{N}\right)^{1/3}\left(\frac{1}{\xi_{E}}\right)\left(\frac{0.1{\rm eV}}{m_{\nu}}\right). (55)

These equations match Eq. (45).

The stable configuration of a cluster corresponds to the hydrostatic equilibrium 𝒫T=𝒫Yuk{\cal P}_{T}=-{\cal P}_{\rm Yuk}. In the non-relativistic case, 𝒫T=(2/3)ρT{\cal P}_{T}=(2/3)\rho_{T} and therefore from Eq. (50) we have EK=3/2V𝒫TE_{K}=3/2V{\cal P}_{T}. So, 𝒫T=(2/3)EK/V{\cal P}_{T}=(2/3)E_{K}/V. The Yukawa pressure equals 𝒫Yuk=EV/V{\cal P}_{\rm Yuk}=E_{V}/V. From this we find that the hydrostatic equilibrium corresponds to 2/3EK=EV2/3E_{K}=E_{V} or ξE=3/2\xi_{E}=3/2.

The ratio of the degenerate ν\nu-cluster radius in Eq. (45), denoted by RdegR_{\rm deg}, to the radius RTR_{T}, equals

RdegRT=0.346ξE=0.52,\frac{R_{\rm deg}}{R_{T}}=0.346\xi_{E}=0.52, (56)

where the last equality is for the hydrostatic equilibrium. The ratio does not depend on other parameters. So, if the evolution proceeds via the formation of a cluster in hydrostatic equilibrium, achieving final configuration require that the radius decreases by a factor of two. Our consideration of the thermal case is oversimplified: applying the Hydrostatic equilibrium locally leads to a distribution of neutrinos in a cluster with a larger radius.

Therefore, the density of a non-degenerate cluster equals

n=mν39103π6I35I53y6N2ξE3,n=m_{\nu}^{3}\frac{9\cdot 10^{-3}}{\pi^{6}}\leavevmode\nobreak\ \frac{I_{3}^{5}}{I_{5}^{3}}\leavevmode\nobreak\ y^{6}N^{2}\xi_{E}^{3},

and numerically

n=6.2106cm3(y107)6(N61022)2ξE3(mν0.1eV)3.n=6.2\cdot 10^{6}\leavevmode\nobreak\ {\rm cm}^{-3}\left(\frac{y}{10^{-7}}\right)^{6}\left(\frac{N}{6\cdot 10^{22}}\right)^{2}\xi_{E}^{3}\left(\frac{m_{\nu}}{0.1{\rm eV}}\right)^{3}.

V Relativistic regime

V.1 Equations for degenerate neutrino cluster

For a sufficiently large NN, neutrinos become relativistic (pFm~νp_{F}\gg\tilde{m}_{\nu}) in the center of cluster, while near the surface they are almost at rest (pF=0)(p_{F}=0). For relativistic neutrinos the Yukawa potential is suppressed by the factor m~ν/E𝐩\tilde{m}_{\nu}/E_{\mathbf{p}} in Eq. (15). Due to this suppression, the pressure of degenerate gas is able to equilibrate the Yukawa attraction for arbitrarily large NN, in contrast to the gravity case.

The system is described by Eq. (15) and relativistic generalization of equilibrium condition between the Yukawa attraction and the degenerate neutrino gas repulsion, Eq. (39). To avoid potential ambiguities in the definition of forces in the relativistic regime, instead of FYukF_{{\rm Yuk}} and FdegF_{{\rm deg}} we will use the derivatives dϕ/drd\phi/dr and dpF/drdp_{F}/dr which represent these two forces and are well-defined quantities in the relativistic case. The equilibrium condition can be derived from the equations of motion for the fields (see Appendix E) giving

y(mν+yϕ)dϕdr=pFdpFdr,y(m_{\nu}+y\phi)\frac{d\phi}{dr}=-p_{F}\frac{dp_{F}}{dr}, (57)

(r<Rr<R) which is the relativistic analogy of equality FYuk=FdegF_{{\rm Yuk}}=F_{{\rm deg}}. In turn, Eq. (15) for the scalar field can be rewritten as

[2mϕ2]yϕ=y2n~,\left[\nabla^{2}-m_{\phi}^{2}\right]y\phi=y^{2}\tilde{n}, (58)

where

n~=m~ν2π20pFp2dpm~ν2+p2=m~ν34π2[pFm~ν1+pF2/m~ν2tanh1(pFpF2+m~ν2)]\tilde{n}=\frac{\tilde{m}_{\nu}}{2\pi^{2}}\int_{0}^{p_{F}}\frac{p^{2}dp}{\sqrt{\tilde{m}_{\nu}^{2}+p^{2}}}=\frac{\tilde{m}_{\nu}^{3}}{4\pi^{2}}\left[\frac{p_{F}}{\tilde{m}_{\nu}}\sqrt{1+p_{F}^{2}/\tilde{m}_{\nu}^{2}}-\tanh^{-1}\left(\frac{p_{F}}{\sqrt{p_{F}^{2}+\tilde{m}_{\nu}^{2}}}\right)\right] (59)

is the effective neutrino density and m~ν\tilde{m}_{\nu} is introduced in Eq. (9). Let us underline that since the effective neutrino mass is given by m~ν\tilde{m}_{\nu}, the relativistic regime is determined by the condition pF/m~ν1p_{F}/\tilde{m}_{\nu}\gg 1 which can differ substantially from pF/mν1p_{F}/m_{\nu}\gg 1.

Using Eq. (59), we can rewrite the system of Eqs. (57) and (58) in terms of m~ν\tilde{m}_{\nu} and pFp_{F}:

[2mϕ2](m~νmν)\displaystyle\left[\nabla^{2}-m_{\phi}^{2}\right](\tilde{m}_{\nu}-m_{\nu}) =y2m~ν34π2[pFm~ν1+pF2/m~ν2tanh1(pFpF2+m~ν2)],\displaystyle=y^{2}\frac{\tilde{m}_{\nu}^{3}}{4\pi^{2}}\left[\frac{p_{F}}{\tilde{m}_{\nu}}\sqrt{1+p_{F}^{2}/\tilde{m}_{\nu}^{2}}-\tanh^{-1}\left(\frac{p_{F}}{\sqrt{p_{F}^{2}+\tilde{m}_{\nu}^{2}}}\right)\right], (60)
m~νdm~νdr\displaystyle\tilde{m}_{\nu}\frac{d\tilde{m}_{\nu}}{dr} =pFdpFdr(r<R).\displaystyle=-p_{F}\frac{dp_{F}}{dr}\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ (r<R). (61)

Eqs. (60) and (61) is a system of differential equations for pFp_{F} and m~ν\tilde{m}_{\nu} with the boundary conditions in the center:

m~ν(r=0)=m~0,pF(r=0)=pF0.\tilde{m}_{\nu}(r=0)=\tilde{m}_{0},\,\,\,\,\,p_{F}(r=0)=p_{F0}. (62)

Eq. (61) can be integrated from r=0r=0 to a given rr:

m~ν2=pF2+m~02+pF02.\tilde{m}_{\nu}^{2}=-p_{F}^{2}+\tilde{m}_{0}^{2}+p_{F0}^{2}.

Inserting m~ν\tilde{m}_{\nu} from this equation into Eq. (60) we obtain a single equation for pFp_{F}, which can be solved numerically444 See Appendix G. Our numerical code is available at https://github.com/xunjiexu/neutrino-cluster. and using n(r)=pF3(r)/6π2n(r)=p_{F}^{3}(r)/6\pi^{2} (which still holds for relativistic neutrinos) to compute the number density.

In the differential equations, the evolution of m~ν\tilde{m}_{\nu} represents the evolution of the scalar field while pFp_{F} determines the neutrino density. Therefore the boundary condition for m~ν\tilde{m}_{\nu} is essentially the condition for the field ϕ\phi. Note that pF0/m~0p_{F0}/\tilde{m}_{0} is the measure of relativistic character of the system. Hence pF0/m~0p_{F0}/\tilde{m}_{0} is used as the main input condition which determines the rest, while m~0\tilde{m}_{0} is determined by a self-consistent condition below.

As in the case of the non-relativistic Lane-Emden equation, at certain r=Rr=R, the Fermi momentum pFp_{F} (and hence the density) drops down to zero:

pF(r=R)=0,p_{F}(r=R)=0,\,\,

and this defines the radius of a cluster. Since the field ϕ\phi extends beyond a star radius, r>Rr>R, the effective mass m~ν\tilde{m}_{\nu} differs from mνm_{\nu} at r>Rr>R. But it should coincide with the vacuum mass mνm_{\nu} at rr\rightarrow\infty. To obtain a self-consistent solution, we need to tune m~0\tilde{m}_{0} to match m~ν(r)\tilde{m}_{\nu}(r\rightarrow\infty) with mνm_{\nu}, i.e. m~0\tilde{m}_{0} is not a free parameter but a parameter determined by the neutrino mass at rr\rightarrow\infty.

The system of Eq. (60) and (61) is similar to that for nuggets in Gresham et al. (2017). Indeed, differentiating Eq. (5) of Gresham et al. (2017) with respect to rr gives exactly Eq. (61). Eq. (60) corresponds (up to a factor of 1/π1/\pi) to Eq. (6) in Gresham et al. (2017). However, the boundary conditions (7) there differs from our boundary conditions determined by Eq. (62) and by the aforementioned matching of m~ν(r)\tilde{m}_{\nu}(r\rightarrow\infty). The boundary condition in the relativistic regime together with the matching is non-trivial as already noticed in Stephenson et al. (1998). This condition can affect the existence of solutions.

Eqs. (60) and (61) can be reduced to the Lane-Emden equation in the non-relativistic limit. Indeed, in this limit n~n\tilde{n}\approx n and for mϕ=0m_{\phi}=0 we obtain from (60)

1r2ddr[r2d(yϕ)dr]=y2n,\frac{1}{r^{2}}\frac{d}{dr}\left[r^{2}\frac{d(y\phi)}{dr}\right]=-y^{2}n, (63)

which is the spherically symmetric case of (60). From (57), neglecting V=yϕV=y\phi in comparison to mνm_{\nu} we find

d(yϕ)dr=12mνdpF2dr=(6π2)2/32mνdn2/3dr.\frac{d(y\phi)}{dr}=-\frac{1}{2m_{\nu}}\frac{dp_{F}^{2}}{dr}=-\frac{(6\pi^{2})^{2/3}}{2m_{\nu}}\frac{dn^{2/3}}{dr}.

Inserting this expression for the derivative into Eq. (63) gives Eq. (42).

V.2 Solution for massless mediator

Refer to caption
Figure 3: The same as in Fig. 2 but for the relativistic case. Shown are the density (solid) and effective density (dashed) profiles for different values of pF0/m~0p_{F0}/\tilde{m}_{0}.

Following the above setup, we first solve the differential equations (60) and (61) with mϕ=0m_{\phi}=0. In Fig. 3, we show the obtained density n(r)n(r) and the effective density n~(r)\tilde{n}(r) profiles for several values of pF0/m~0p_{F0}/\tilde{m}_{0}. Recall that the effective density is the number density that includes the suppression factor m~ν/E𝐩\tilde{m}_{\nu}/E_{\mathbf{p}}. This factor depends on pFp_{F}, and decreases toward the center. As a result, suppression in the center is stronger and the maximum of n~(r)\tilde{n}(r) occurs at certain point away from r=0r=0. This effect becomes stronger with the increase of pF0/m~0p_{F0}/\tilde{m}_{0}.

Using the obtained density profiles we computed various characteristics of the neutrino clusters for y=107y=10^{-7} and mϕ=0m_{\phi}=0 (see Tab. 1). The correlations of characteristic parameters (RR, NN, pF0p_{F0}) are shown by blue lines in Figs. 4 and 5. For a given value of pF0/m~0p_{F0}/\tilde{m}_{0} (second line in Tab. 1) we determined m~0/mν\tilde{m}_{0}/m_{\nu} (the third line). The total numbers of neutrinos NN (the first line) was computed according to

N=0Rn(r)4πr2𝑑r.N=\int_{0}^{R}n(r)4\pi r^{2}dr.

Using the numbers in the second and third lines we find pF0/mνp_{F0}/m_{\nu} (the 4th line), which in turn, gives the central density of a ν\nu-cluster:

n0=2.1109cm3(pF0mν)3(mν0.1eV)3.n_{0}=2.1\cdot 10^{9}\leavevmode\nobreak\ {\rm cm^{-3}}\leavevmode\nobreak\ \left(\frac{p_{F0}}{m_{\nu}}\right)^{3}\left(\frac{m_{\nu}}{0.1\leavevmode\nobreak\ {\rm eV}}\right)^{3}\,.

As NN increases, the central density n0n_{0} (also pF0p_{F0}) first increases, reaches maximum at Ny3=1.5102Ny^{3}=1.5\cdot 10^{2} or N=1.51023N=1.5\cdot 10^{23} and then it decreases. The maximum corresponds to pF00.6mνp_{F0}\simeq 0.6\leavevmode\nobreak\ m_{\nu}, that is, to the point of transition between the non-relativistic and relativistic regimes. The maximal density is determined by the mass of neutrino only:

nνmax(0)=4.3108cm3(mν0.1eV)3.n_{\nu}^{\max}(0)=4.3\cdot 10^{8}\leavevmode\nobreak\ {\rm cm^{-3}}\left(\frac{m_{\nu}}{0.1{\rm eV}}\right)^{3}.

The corresponding values of radius of a cluster are shown in the last line of Tab. 1. On the contrary, as NN increases, the radius RR first (in non-relativistic range) decreases, reaches the minimum

Rmin0.62km(107y)(0.1eVmν)kmR_{\min}\approx 0.62\leavevmode\nobreak\ {\rm km}\left(\frac{10^{-7}}{y}\right)\left(\frac{0.1\ {\rm eV}}{m_{\nu}}\right){\rm km}\thinspace (64)

at N1.51023N\approx 1.5\cdot 10^{23} and then increases, as shown in Fig. 4. The minimal radius (64) is about 20% smaller than the non-relativistic result in Eq. (43) for pF=0.3mνp_{F}=0.3m_{\nu}, which can still be accurately computed using the non-relativistic approximation according to the difference between the dashed and solid curves in Fig. 3.

Notice that dependence n(r)n(r) substantially differs from n(r)=pF3(r)/6π2n(r)=p_{F}^{3}(r)/6\pi^{2} with pF(r)p_{F}(r) taken from the ansatzes of Wise and Zhang (2015). For some distances rr, the difference is given by a factor of 5 to 6. At the same time there is good agreement of our n(r)n(r) shown in Fig. 3 with that obtained in Gresham et al. (2017).

Table 1: Characteristics of final (degenerate) states of neutrino clusters for y=107y=10^{-7} and mϕ=0m_{\phi}=0.
NN 2.9610212.96\cdot 10^{21} 1.6310221.63\cdot 10^{22} 5.9610225.96\cdot 10^{22} 9.3510239.35\cdot 10^{23} 2.3410242.34\cdot 10^{24}
pF0/m~0p_{F0}/\tilde{m}_{0} 0.10 0.31 0.75 7.0 22
m~0/mν\tilde{m}_{0}/m_{\nu} 0.991 0.922 0.688 0.060 0.014
pF0/mνp_{F0}/m_{\nu} 0.099 0.286 0.561 0.420 0.308
n0n_{0} [cm-3] 2.01062.0\cdot 10^{6} 4.91074.9\cdot 10^{7} 3.71083.7\cdot 10^{8} 1.51081.5\cdot 10^{8} 6.11076.1\cdot 10^{7}
RR [km] 1.25 0.75 0.62 1.46 2.41

The effective neutrino mass at the border of the cluster can be obtained as m~ν(R)=mν+V(R)\tilde{m}_{\nu}(R)=m_{\nu}+V(R) with V(R)V(R) defined in Eq. (13):

m~ν(R)=mνy2mϕRemϕR0Rrn~(r)sinh(mϕr)𝑑r.\tilde{m}_{\nu}(R)=m_{\nu}-\frac{y^{2}}{m_{\phi}R}e^{-m_{\phi}R}\int_{0}^{R}r\tilde{n}(r)\sinh(m_{\phi}r)dr.

Here n~(r)\tilde{n}(r) should be computed according to Eq. (59). For mϕ=0m_{\phi}=0 the expression simplifies to

m~ν(R)=mνy2R0Rn~(r)r2𝑑r.\tilde{m}_{\nu}(R)=m_{\nu}-\frac{y^{2}}{R}\int_{0}^{R}\tilde{n}(r)r^{2}dr.

Dependence of NN and RR on yy can be obtained using rescaling: rr¯=ry,mϕm¯ϕ=mϕ/yr\to\overline{r}=ry,m_{\phi}\to\overline{m}_{\phi}=m_{\phi}/y — see Eq. (151) in Appendix G. The rescaling means that relative effects of the density gradient and mass of ϕ\phi in the equation for m~ν\tilde{m}_{\nu} do not change with yy, if mϕ/ym_{\phi}/y is given at a fixed value. We can write

R(y)=R7(107y),N(y)=N7(107y)3,R(y)=R_{7}\left(\frac{10^{-7}}{y}\right),\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ N(y)=N_{7}\left(\frac{10^{-7}}{y}\right)^{3}, (65)

where R7R_{7} and N7N_{7} are the radius and density at y=107y=10^{-7} given in the Table I. From Eq. (65), we find e.g. that at y=1014y=10^{-14} and parameters in the last column of the Table I, R=2.41012R=2.4\cdot 10^{12} cm and N=2.31045N=2.3\cdot 10^{45}, which corresponds to M=3.31012M=3.3\cdot 10^{12} g. For mϕ=0m_{\phi}=0, the distribution in rr simply dilates with yy, and NN changes correspondingly.

According to Tab. 2, for small NN the field ϕ\phi in a cluster is weak, therefore m~νmν\tilde{m}_{\nu}\approx m_{\nu} and pF0/m~0pF0/mνp_{F0}/\tilde{m}_{0}\approx p_{F0}/m_{\nu}. With increase of NN the field increases, m~0\tilde{m}_{0} decreases, i.e., the medium suppresses the effective neutrino mass. N7=1.51023N_{7}=1.5\cdot 10^{23} is critical value between the non-relativistic and relativistic cases. Further increasing NN, the dependence of RR and n0n_{0} on NN becomes opposite: RR increases, while n0n_{0} decreases.

V.3 Neutrino clusters for nonzero mϕm_{\phi}

The dependence of properties of a ν\nu-cluster on mϕm_{\phi} is rather complicated. With increase of mϕm_{\phi} the radius of scalar interaction becomes smaller and hence the interactions become weaker. Therefore, in general, one expects that the binding effect weakens and the system becomes less compact for a fixed NN. As a result, the radius RR increases with increase of mϕm_{\phi}, which indeed is realized in the non-relativistic case.

Refer to caption
Figure 4: Dependence of the radius RR of a neutrino degenerate cluster in the unit of ymνym_{\nu} on Ny3Ny^{3} for different values of mϕm_{\phi}.

In Fig. 4 we show the dependence of the radius RR on NN for different values of mϕm_{\phi}. For a given mϕm_{\phi} the dependence has a minimum RminR_{\min}. As in the massless case, with increase of NN the radius first decreases, reaches minimum and then increases in the relativistic regime. This behavior is explained by the fact that in the relativistic case the attractive force is suppressed by m~ν/E\tilde{m}_{\nu}/E, and with increase of NN the effective mass m~ν\tilde{m}_{\nu} decreases.

The main features of dependence of RR on NN are discussed below.

(i) Existence of a lower bound on NN. At certain NN the lines become vertical. The bound is absent for mϕ=0m_{\phi}=0. For each nonzero mϕm_{\phi}, there is a lower bound, NminN_{\min}, below which it is impossible to make the neutrinos confined. Since the potential energy (attraction) increases with NN, to support bound system one should have large enough number of neutrinos. NminN_{\min} increases with increase of mϕm_{\phi}, that is, with decrease of radius of interactions of individual neutrinos 1/mϕ1/m_{\phi}. The smaller the radius, the larger NN should be to keep the system bounded.

The dependence of RR on NN in Fig. 4 differs from that in Fig. 2 of Gresham et al. (2017). In the region of NN around minimal radius and above that for mϕ=0m_{\phi}=0 radius in Fig. 4 is larger than the radius in Wise and Zhang (2015) and Gresham et al. (2017) and the difference increases with decrease of NN. The difference is due to that Gresham et al. (2017) used an analytical solution for ϕ\phi which is only valid when n~\tilde{n} is a constant.

Furthermore, the curves R=R(N,mϕ)R=R(N,m_{\phi}) of Gresham et al. (2017) do not extend far below NN that correspods to the minimum of RR, since the number NN becomes small. Indeed, in Gresham et al. (2017) the constant αϕ(0.010.1)\alpha_{\phi}\sim(0.01-0.1) corresponds to y3=(0.0451.4)y^{3}=(0.045-1.4). For N102N\geq 10^{2} this gives Ny3(4.5140)Ny^{3}\geq(4.5-140). Our results extend to much smaller values of Ny3Ny^{3}, where important feature of existence of minimal value of NN for mϕ0m_{\phi}\neq 0 is realized.

The dependence of RR on mϕm_{\phi} for fixed NN is also similar here to those in Gresham et al. (2017). In the relativistic branch the radius decreases with the increase of mϕm_{\phi}.

For mϕ=0m_{\phi}=0 (blue line) the dependence of RR on NN is well reproduced by Eq. (45) in the non-relativistic case. Notice that in this case the scale of RR is related to the neutrino mass.

(ii) Existence of a minimal radius for a given mϕm_{\phi}. The minimum corresponds to pF0m~νp_{F0}\sim\tilde{m}_{\nu} (i.e. to the effective mass of neutrinos which can be much smaller than the vacuum mass). This minimal radius, RminR_{\min}, in the mϕ=0m_{\phi}=0 case is determined by Eq. (64). For nonzero mϕm_{\phi}, it becomes larger because nonzero mϕm_{\phi} suppresses the Yukawa potential exponentially.

With increase of mϕm_{\phi} the minimum RminR_{\min} shifts to larger NN. The value RminR_{\min} itself changes weakly. For large mϕm_{\phi} the minimum of R(N)R(N) does not correspond to m~νpF0\tilde{m}_{\nu}\sim p_{F0}, but still occurs close to transition from NR to UR regimes. Because of the shift of the curves for fixed NN in the NR regime the radius increases with increase of mϕm_{\phi}, while in the UR case it decreases. The shift can be interpreted in the following way: For fixed RR the volume per neutrino is determined by mϕm_{\phi}: V1(1/mϕ)κV_{1}\propto(1/m_{\phi})^{\kappa} (κ3\kappa\sim 3). Therefore with increase of mϕm_{\phi}, the volume per neutrino decreases and the total number of NN in a cluster increases.

There is maximal value of mϕm_{\phi} for which solution with finite radius exists. With approaching of mϕm_{\phi} to this value the curve further shifts to larger NN and minimum moves up. For mϕ>mϕmaxm_{\phi}>m_{\phi}^{\max} the radius of interaction becomes too short to bound the neutrino system and the system disintegrates. The upper bound on mϕm_{\phi} corresponds to the lower bound on strength of interaction:

Sϕy2mν2mϕ2Sϕmin=70,S_{\phi}\equiv\frac{y^{2}m_{\nu}^{2}}{m_{\phi}^{2}}\gtrsim S_{\phi}^{\min}=70\,, (66)

which can be compared with the bound Gϕ>3.27G_{\phi}>3.27 found in Stephenson et al. (1998) from the condition of existence of minimum of total energy per neutrino for infinite medium. Notice that our bound is obtained from condition of stability of finite configuration.

Let us compare the radius RR with the radius of interactions Rϕ=1/mϕR_{\phi}=1/m_{\phi}. For small mϕm_{\phi} there is the range of values of NN in which R<1/mϕR<1/m_{\phi}. It is possible to obtain R=RϕR=R_{\phi} at two different values of NN. For mϕ=0.03ymνm_{\phi}=0.03ym_{\nu} the equality R=RϕR=R_{\phi} is fulfilled at definite value N=102/y3N=10^{2}/y^{3}. For mϕ>0.03ymνm_{\phi}>0.03ym_{\nu}, we have R>RϕR>R_{\phi} for all NN and the difference between them increases. RR can be orders of magnitude larger than RϕR_{\phi} and therefore in general 1/mϕ1/m_{\phi} does not determine the size of the cloud and fragmentation scale.

We denote by XX and YY the values at horizontal and vertical axes of Fig. 4 correspondingly:

XNy3andYRymν.X\equiv Ny^{3}\leavevmode\nobreak\ \leavevmode\nobreak\ {\rm and}\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ Y\equiv Rym_{\nu}.

The coordinate YY can be written as Y=R/RϕSϕ1/2Y=R/R_{\phi}S_{\phi}^{-1/2}. Therefore the radius of a cluster over the radius of interactions equals

RRϕ=YSϕ1/2.\frac{R}{R_{\phi}}=YS_{\phi}^{-1/2}. (67)

For Sϕ1/2=0.01S_{\phi}^{-1/2}=0.01 the radius (in units ymνym_{\nu}) is in the interval Y=(30180)Y=(30-180) (see Fig. 4) which gives according to Eq.(67) R/Rϕ=0.31.8R/R_{\phi}=0.3-1.8. For Sϕ1/2=0.02S_{\phi}^{-1/2}=0.02 the corresponding numbers are Y=(30170)Y=(30-170) and R/Rϕ=(0.63.4)R/R_{\phi}=(0.6-3.4). For Sϕ1/2=0.1S_{\phi}^{-1/2}=0.1 we have Y=(55110)Y=(55-110) and R/Rϕ=511R/R_{\phi}=5-11. Thus, with decrease of strength the radius increases and becomes substantially larger than RϕR_{\phi}. This regime was observed in Gresham et al. (2017) and called “saturation”. Still 1/mϕ1/m_{\phi} determines the scale of sizes of ν\nu-clusters within an order of magnitude.

Refer to caption
Figure 5: Dependence of the Fermi momentum in the center of ν\nu-cluster, pF0p_{F0}, on Ny3Ny^{3} for different values of mϕm_{\phi} (mϕ/ymνm_{\phi}/ym_{\nu}).

In Fig. 5 we show the dependence of pF0p_{F0} (and consequently, the central density) on NN for different values of mϕm_{\phi}. For fixed mϕm_{\phi}, pF0p_{F0} as a function of NN has a maximum. With increase of mϕm_{\phi}, this maximum shifts to larger NN and increases. However, this increase slows down and is restricted by pF0/mν0.9p_{F0}/m_{\nu}\lesssim 0.9. Explicitly, for this value

n0max=1.5109cm3(mν0.1eV)3.n_{0}^{\max}=1.5\cdot 10^{9}{\rm cm^{-3}}\left(\frac{m_{\nu}}{0.1{\rm eV}}\right)^{3}. (68)

With increase of mϕm_{\phi} the central density decreases at small NN (the NR regime) and it increases at larger NN (UR regime). This anti-correlates with change of radius (for fixed NN) in Fig. 4. The maximal density (corresponding to the maximal pF0p_{F0}) increases with mϕm_{\phi} and shifts to larger NN. The anticorrelation is due to the relation

Nb4π3R3nF0=b29πR3pF03,N\sim b\frac{4\pi}{3}R^{3}n_{F0}=b\frac{2}{9}\pi R^{3}p_{F0}^{3}, (69)

where b0.20.5b\approx 0.2-0.5 is smaller than 1, since the density distribution is not uniform and the density is substantially smaller than nF0n_{F0} in peripheral regions.

Thus, the overall change of characteristics of cluster (for fixed NN) with mϕm_{\phi} is the following: In the NR case central density decreases and radius increases (cluster becomes less compact). In the UR case, the radius varies non-monotonically with mϕm_{\phi}, as shown by the relativistic branches of the curves in Fig. 4. These characteristics could be important for cluster formation processes.

For fixed mϕm_{\phi} the bound system exists for NN above certain minimal value. With the increase of NN the central density (and pF0p_{F0}) increases quickly and therefore the radius of cluster decreases. Then the density reaches its maximum and with further increase of NN the density decreases. Correspondingly the radius of cluster increases.

Notice that strengths of interactions in the case of two neutrino bound states λ\lambda [see Eq. (19)] and for the bound neutrino system [Eq. (66)] have different functional dependence on the parameters. From Eqs. (19) and (66) we obtain the following relation between the two strengths:

λ2=(y8π)2Sϕ.\lambda^{2}=\left(\frac{y}{8\pi}\right)^{2}S_{\phi}.

For y107y\leq 10^{-7} and λ20.7\lambda^{2}\geq 0.7 required to have 2ν2\nu-bound state we find Sϕ>6.31016S_{\phi}>6.3\cdot 10^{16}. That is, if 2ν2\nu-bound states exist also ν\nu-clusters should exist. The opposite does not work: For Sϕ102S_{\phi}\sim 10^{2} we have λ1.61015\lambda\sim 1.6\cdot 10^{-15}, and so the 2ν2\nu bound states are not possible. Both the bound states and bound systems exist for very small masses of scalars.

V.4 Neutrino bound systems for given yy and mϕm_{\phi}

In the previous sections we focused on properties of final stable configurations of neutrino clusters. Let us comment on implications of the obtained results.

From the particle physics perspective the input parameters are yy and mϕm_{\phi}, and the questions are whether bound systems for given yy and mϕm_{\phi} exist, and what the characteristic parameters of these systems are. The parameters yy and mϕm_{\phi} determine the interaction strength SϕS_{\phi} defined in Eq. (66). The main condition for existence of bound system is that the strength SϕS_{\phi} is large enough, satisfying the lower bound (66) or Sϕ1/20.12S_{\phi}^{-1/2}\leq 0.12. For a given yy, the condition on the strength gives an upper bound on mϕm_{\phi}. So, essentially yy and SϕS_{\phi} determine all other characteristics.

Let us consider properties of clusters at the critical (minimal) value of strength. According to Fig. 4, SϕminS_{\phi}^{\min} is realized at

Xcr104,Ycr102X^{\rm cr}\approx 10^{4},\,\,\,Y^{\rm cr}\approx 10^{2}

with very small spread. As follows from Fig. 5, it corresponds to maximal pF0/mν=0.86p_{F0}/m_{\nu}=0.86 and therefore maximal central density of the cluster. With these specific values of XX and YY, the parameters of cluster are determined by yy. Taking y=107y=10^{-7}, we obtain

mϕcr108mν=109eV,m_{\phi}^{\rm cr}\approx 10^{-8}m_{\nu}=10^{-9}\,{\rm eV},

and

Ncr=Xcry31025,Rcr=2104cmYcry12105cm.N^{\rm cr}=X^{\rm cr}y^{-3}\approx 10^{25},\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ R^{\rm cr}=2\cdot 10^{-4}{\rm cm}\leavevmode\nobreak\ Y^{\rm cr}y^{-1}\approx 2\cdot 10^{5}{\rm cm}.

If mϕ<mϕcrm_{\phi}<m_{\phi}^{\rm cr} (for fixed yy), the strength increases and according to Fig. 4, the ranges of XX and YY expand. Correspondingly,

N=Xy3=1025X/Xcr,N=Xy^{-3}=10^{25}X/X^{\rm cr},

and therefore clusters with much smaller number of neutrinos are possible. E.g. for Sϕ=104S_{\phi}=10^{4}, XX can be as small as 0.2, and therefore N=21020N=2\cdot 10^{20}. XX becomes a free parameter in the interval 0.21040.2-10^{4}. The allowed range of YY expands in both directions: YY can be as small as 0.3Ycr0.3Y^{\rm cr}, and therefore the radius reduces down to 0.70.7 km. The maximal value of YY can be much bigger than YcrY^{\rm cr} and it increases with increase of strength (decrease of mϕm_{\phi}).

R=2104cmYy1=2105cmY/Ycr.R=2\cdot 10^{-4}{\rm cm}\leavevmode\nobreak\ Yy^{-1}=2\cdot 10^{5}{\rm cm}\leavevmode\nobreak\ Y/Y^{\rm cr}.

For Sϕ=104S_{\phi}=10^{4} (orange line in Fig. 4) Y=(30380)Y=(30-380) and correspondingly, R=(0.67.6)R=(0.6-7.6) km. If also yy decreases and mϕm_{\phi} decreases, so that SϕS_{\phi} does not change, the line Y=Y(X)Y=Y(X) is the same, but parameters of cluster increase as Ny3N\propto y^{-3}, Ry1R\propto y^{-1}.

For fixed yy and mϕm_{\phi} stable bound systems are situated along the corresponding lines Sϕ=constS_{\phi}={\rm const} in Figs. 4 and 5. This, in turn, fixes the interval of XX and YY (hence NN and RR):

N1y3[Xmin,Xmax],R1ymν[Ymin,Ymax].N\in\frac{1}{y^{3}}\left[X_{{\rm min}},\,X_{{\rm max}}\right]\thinspace,\,\,\,\,\,R\in\frac{1}{ym_{\nu}}[Y_{{\rm min}},\,Y_{{\rm max}}]\thinspace. (70)

As an example, for y=1020y=10^{-20} and mϕ=1023m_{\phi}=10^{-23} eV we obtain Sϕ1/2=0.01S_{\phi}^{-1/2}=0.01 (orange line), and according to Fig. 4 we have X[0.2, 5103]X\in[0.2,\ 5\cdot 10^{3}]. For chosen yy, this gives N[21059, 51063]N\in[2\cdot 10^{59},\ 5\cdot 10^{63}]. Depending on NN, the central density changes (see Fig. 5) following the relation (69). The interval Y[30, 370]Y\in[30,\ 370] (Fig. 4) corresponds to the interval of values R[61017, 7.41019]R\in[6\cdot 10^{17},\ 7.4\cdot 10^{19}] cm and R/Rϕ[0.3, 3]R/R_{\phi}\in[0.3,\ 3]. With increase of yy for fixed strength, NN and RR change as in Eq. (70). With decrease of strength, the intervals of XX and YY shrink. Thus for Sϕ1/2=0.1S_{\phi}^{-1/2}=0.1, we have X[400, 2104]X\in[400,\ 2\cdot 10^{4}] and Y[50, 110]Y\in[50,\ 110]. If y=1020y=10^{-20} and mϕ=1022m_{\phi}=10^{-22} eV, we obtain N[41062, 21064]N\in[4\cdot 10^{62},\ 2\cdot 10^{64}] and R[1018, 2.31018]R\in[10^{18},\ 2.3\cdot 10^{18}] cm.

From the cosmological perspective the key input parameter is NN. It is this quantity that determines in the nearly uniform background evolution and formation of clusters. If NN is conserved (no merging of clusters, no particle loss, etc.), this quantity determines the final radius and density of the cluster. From Eq. (45) we obtain

R=1.8km(107y)2(1021N)1/3,R=1.8\leavevmode\nobreak\ {\rm km}\leavevmode\nobreak\ \left(\frac{10^{-7}}{y}\right)^{2}\left(\frac{10^{21}}{N}\right)^{1/3},

and from Eqs. (44) and (46):

n0=2.25108cm3(y3N)2.n_{0}=2.25\cdot 10^{8}\leavevmode\nobreak\ {\rm cm^{-3}}(y^{3}N)^{2}.

These relations are valid for mϕ1/Rm_{\phi}\ll 1/R. With decrease of yy (and fixed NN), RR increases as 1/y1/y and the density decreases as y3/2y^{3/2}.

VI Formation of neutrino clusters

Let us consider the formation of neutrino clusters from the uniform background of relic neutrinos in the course of cooling and expansion of the Universe.

We find that for the allowed values of yy and mνm_{\nu}, the typical time of cooling via ϕ\phi bremsstrahlung and via annihilation νν¯ϕϕ\nu\bar{\nu}\rightarrow\phi\phi are much larger than the age of the Universe (see Appendices I and J). Therefore the formation of neutrino clusters differs from the formation of usual stars.555 Notice that the formation of neutrino clusters is very different from the formation of nuggets of ADM Gresham et al. (2018). The latter proceeds via fusion of particles of DM and requires C-asymmetry to avoid complete annihilation. As we will show in subsections VI.3 and VI.4, for the interaction strength above a certain critical value it has a character of phase transition at which the kinetic energy is transformed into the increase of the effective neutrino mass. The mechanism was first suggested in Stephenson et al. (1998) and here we explore it further.

The phase transition leads to fragmentation of the cosmic neutrino background. Here we present some simplified arguments and analytic estimations. A complete solution of the formation problem would require numerical simulation of evolution of the ν\nu background. Such a simulation is beyond the scope of this paper.

VI.1 Maximal density and fragmentation of the ν\nu background

The key parameter that determines formation of ν\nu-clusters is the central density given by Eq. (68). According to Fig. 5, there is a maximal value of pF0p_{F0}, and consequently, a maximal value of the density n0maxn_{0}^{\rm max} which is determined by the mass of neutrino in Eq. (68). After the fragmentation of a uniform relic ν\nu-background (see below), the central density of a cluster stopped decreasing. Consequently, at the epoch of fragmentation the density was

nfragn0max.n_{\rm frag}\leq n_{0}^{\max}.

Thus, we can take

nfrag109cm3.n_{\rm frag}\leq 10^{9}\leavevmode\nobreak\ {\rm cm^{-3}}.

In turn, this density determines the epoch of fragmentation:

(zf+1)=(nfragnrel)1/3=215,(z_{f}+1)=\left(\frac{n_{\rm frag}}{n_{\rm rel}}\right)^{1/3}=215, (71)

where nrel=113n_{\rm rel}=113 is the number density of one neutrino mass state that would be in the present epoch in the absence of clustering. That is, formation of ν\nu-clusters starts after the epoch of recombination (z=1100z=1100, T=0.1T=0.1 eV). At this epoch (zf+1=215z_{f}+1=215) the neutrino average momentum was p0.02p\sim 0.02 eV. So, fragmentation could start when pmνp\sim m_{\nu}, which is not accidental since n0maxn_{0}^{\max} is determined by mνm_{\nu} only.

VI.2 Evolution of the effective neutrino mass

Dependence of the effective mass m~ν\tilde{m}_{\nu} on TT plays crucial role in the formation of neutrino bound systems. If neutrinos are the only source of the field ϕ\phi, for the allowed coupling constant ϕ\phi cannot be thermalized in the early Universe and the amount of ϕ\phi particles is negligible. Evolution of the scalar field in expanding universe is described by Eq. (8) with an additional expansion term 3H(t)dϕ/dt3H(t)d\phi/dt, where the Hubble constant depends on time as H(t)=1/2tH(t)=1/2t in the radiation-dominated epoch and as H(t)=2/3tH(t)=2/3t in the matter-dominated epoch.

Notice that in general mϕ0m_{\phi}\neq 0 regularizes the procedure of computations. Otherwise, it is regularized by the Hubble expansion rate HH. Indeed, from the equation of motion (8) for uniform distribution, assuming ϕ˙=aHϕ\dot{\phi}=aH\phi, where aa is the scale factor in the Friedmann-Robertson-Walker (FRW) metric, we obtain

ϕ=ya2H2+mϕ2ν¯ν,\phi=-\frac{y}{a^{2}H^{2}+m_{\phi}^{2}}\langle\bar{\nu}\nu\rangle, (72)

and ν¯ν\langle\bar{\nu}\nu\rangle is given in (10). Thus, the expansion term regularizes the solution for mϕ0m_{\phi}\rightarrow 0. The energy density in the field:

ρϕ=y22mϕ2(a2H2+mϕ2)2|ν¯ν|2.\rho_{\phi}=\frac{y^{2}}{2}\frac{m_{\phi}^{2}}{(a^{2}H^{2}+m_{\phi}^{2})^{2}}|\langle\bar{\nu}\nu\rangle|^{2}. (73)

In what follows we assume that HmϕH\ll m_{\phi}, so that expansion can be treated as slow change of density and temperature in the solution of equation without expansion.

Neglecting the Hubble expansion term, Eq. (72) gives for uniform medium

ϕ=ymϕ2ν¯ν.\phi=-\frac{y}{m_{\phi}^{2}}\langle\bar{\nu}\nu\rangle. (74)

From Eq. (74), we obtain the equation for the effective mass of neutrino:

m~ν=mνy2mϕ2ν¯ν,\tilde{m}_{\nu}=m_{\nu}-\frac{y^{2}}{m_{\phi}^{2}}\langle\bar{\nu}\nu\rangle,

where [see Eq. (10)]

ν¯ν=m~ν2π20dpp2p2+m~ν2f(p,T),\langle\bar{\nu}\nu\rangle=\frac{\tilde{m}_{\nu}}{2\pi^{2}}\int_{0}^{\infty}\frac{dp\leavevmode\nobreak\ p^{2}}{\sqrt{p^{2}+\tilde{m}_{\nu}^{2}}}f(p,T),

and f(p,T)f(p,T) is the distribution of neutrinos (thermal or degenerate666Notice that results for the degenerate case can be obtained by substituting I3T3pF33I_{3}T^{3}\rightarrow\frac{p_{F}^{3}}{3} in the results for the thermal case.). Explicitly the equation for m~ν\tilde{m}_{\nu} can be written as

m~ν=mνy2mϕ2m~ν2π20dpp2p2+m~ν2f(p,T).\tilde{m}_{\nu}=m_{\nu}-\frac{y^{2}}{m_{\phi}^{2}}\frac{\tilde{m}_{\nu}}{2\pi^{2}}\int_{0}^{\infty}\frac{dp\leavevmode\nobreak\ p^{2}}{\sqrt{p^{2}+\tilde{m}_{\nu}^{2}}}f(p,T).

Here we use the thermal distribution with T0T\neq 0, which is relevant for T>mνT>m_{\nu}, while in Stephenson et al. (1998) the distribution of degenerate Fermi gas was used. Qualitatively, the results are similar.

Refer to caption
Figure 6: Dependence of the effective neutrino mass m~ν\tilde{m}_{\nu} on T/mνT/m_{\nu} for different values of mϕm_{\phi}.

In Fig. 6, we show dependence of m~ν\tilde{m}_{\nu} on TT for different values of strength of interaction. In the ultra-relativistic limit, Tm~T\gg\tilde{m}, we obtain

m~ν=mνy2m~νT2mϕ2I22π2,\tilde{m}_{\nu}=m_{\nu}-\frac{y^{2}\tilde{m}_{\nu}T^{2}}{m_{\phi}^{2}}\frac{I_{2}}{2\pi^{2}}, (75)

where in general, the integrals InI_{n} have been introduced in Eq. (48). From Eq. (75) we find

m~ν=mν1+y2T2I22π2mϕ2mν2π2mϕ2y2T2I2=mν24mϕ2y2T2=mν24Sϕ(mνT)2m~νrel.\tilde{m}_{\nu}=\frac{m_{\nu}}{1+\frac{y^{2}T^{2}I_{2}}{2\pi^{2}m_{\phi}^{2}}}\approx m_{\nu}\frac{2\pi^{2}m_{\phi}^{2}}{y^{2}T^{2}I_{2}}=m_{\nu}\frac{24m_{\phi}^{2}}{y^{2}T^{2}}=m_{\nu}\frac{24}{S_{\phi}}\left(\frac{m_{\nu}}{T}\right)^{2}\equiv\tilde{m}_{\nu}^{\rm rel}. (76)

That is, the effective mass m~νT2\tilde{m}_{\nu}\propto T^{-2} decreases as TT increases and becomes negligible in the early Universe.

In the non-relativistic case, Tm~νT\ll\tilde{m}_{\nu}, the solutions is

m~ν=mνy2T3mϕ2I32π2m~νnr.\tilde{m}_{\nu}=m_{\nu}-\frac{y^{2}T^{3}}{m_{\phi}^{2}}\frac{I_{3}}{2\pi^{2}}\equiv\tilde{m}_{\nu}^{\rm nr}. (77)

That is, as T0T\to 0, m~νmν\tilde{m}_{\nu}\rightarrow m_{\nu}. These analytic results agree well with results of numerical computations shown in Fig. 6.

Note that the relativistic regime is determined by m~ν\tilde{m}_{\nu} and not mνm_{\nu}: T>m~νT>\tilde{m}_{\nu}. The effective mass m~ν\tilde{m}_{\nu} increases as 1/T21/T^{2} till Tm~νT\approx\tilde{m}_{\nu}. According to Eq. (76) this equality gives

Trelmν=(24Sϕ)1/3.\frac{T_{\rm rel}}{m_{\nu}}=\left(\frac{24}{S_{\phi}}\right)^{1/3}. (78)

Below this temperature the growth of m~\tilde{m} is first faster and then slower than 1/T21/T^{2}. The change of derivative occurs at maximum of the relative energy in the scalar field (see below). m~(T)\tilde{m}(T) converges to the non-relativistic regime at m~νrel(T)=mν\tilde{m}_{\nu}^{\rm rel}(T)=m_{\nu}. This gives, according to Eq. (76),

Tnrmν=(24Sϕ)1/2.\frac{T_{\rm nr}}{m_{\nu}}=\left(\frac{24}{S_{\phi}}\right)^{1/2}.

VI.3 Energy of the ν\nu-ϕ\phi system and the dip

Let us compute the energy density in the scalar field, ρϕ\rho_{\phi} , in neutrinos, ρν\rho_{\nu}, and the total energy density defined as

ρtot=ρν+ρϕ\rho_{\rm tot}=\rho_{\nu}+\rho_{\phi} (79)

as a function of TT. Expansion of the Universe can be accounted for by considering the average energy per neutrino:

ϵiρinν,i={ν,ϕ,tot},\epsilon_{i}\equiv\frac{\rho_{i}}{n_{\nu}},\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ i=\{\nu,\leavevmode\nobreak\ \leavevmode\nobreak\ \phi,\leavevmode\nobreak\ \leavevmode\nobreak\ {\rm tot}\}, (80)

where the number density of neutrinos equals

nν=12π20𝑑pp2f(p,T)=I32π2T3.n_{\nu}=\frac{1}{2\pi^{2}}\int_{0}^{\infty}dp\leavevmode\nobreak\ p^{2}f(p,T)=\frac{I_{3}}{2\pi^{2}}T^{3}.

The energy density in the scalar field is given by

ρϕ=12mϕ2ϕ2.\rho_{\phi}=\frac{1}{2}m_{\phi}^{2}\phi^{2}. (81)

Here ϕ\phi can be expressed via the effective mass, ϕ=(m~νmν)/y\phi=(\tilde{m}_{\nu}-m_{\nu})/y, and therefore

ρϕ(m~ν)=mϕ22y2(m~νmν)2.\rho_{\phi}(\tilde{m}_{\nu})=\frac{m_{\phi}^{2}}{2y^{2}}(\tilde{m}_{\nu}-m_{\nu})^{2}. (82)

So, ρϕ\rho_{\phi} is directly determined by the deviation of the effective mass of neutrino from the vacuum mass. The energy per neutrino equals

ϵϕ=mνπ2I3Sϕ(m~νmν1)2(mνT)3.\epsilon_{\phi}=m_{\nu}\frac{\pi^{2}}{I_{3}S_{\phi}}\left(\frac{\tilde{m}_{\nu}}{m_{\nu}}-1\right)^{2}\left(\frac{m_{\nu}}{T}\right)^{3}. (83)

Using the known form of m~ν(T)\tilde{m}_{\nu}(T), Eq. (75), we compute the energy density ϵϕ\epsilon_{\phi}. In the ultra-relativistic limit, according to Eq. (82), the energy density in ϕ\phi converges to the constant:

ρϕrel=mϕ22y2mν2.\rho_{\phi}^{\rm rel}=\frac{m_{\phi}^{2}}{2y^{2}}m_{\nu}^{2}.

Therefore

ϵϕrel=mνχ,\epsilon_{\phi}^{\rm rel}=\frac{m_{\nu}}{\chi},

where

χSϕI3π2(Tmν)3.\chi\equiv\frac{S_{\phi}I_{3}}{\pi^{2}}\left(\frac{T}{m_{\nu}}\right)^{3}.

In the non-relativistic limit

ρϕnr=18π4y2mϕ2I32T6.\rho_{\phi}^{\rm nr}=\frac{1}{8\pi^{4}}\frac{y^{2}}{m_{\phi}^{2}}I_{3}^{2}T^{6}.

and

ϵϕnr=mνSϕI34π2(Tmν)3=mν4χ.\epsilon_{\phi}^{\rm nr}=m_{\nu}\frac{S_{\phi}I_{3}}{4\pi^{2}}\left(\frac{T}{m_{\nu}}\right)^{3}=\frac{m_{\nu}}{4}\chi. (84)

According to Eq. (84), as the Universe cools down, ϵϕ\epsilon_{\phi} first increases and then decreases. The maximum of ϵϕ\epsilon_{\phi} is achieved when ϵϕrelϵϕnr\epsilon_{\phi}^{\rm rel}\approx\epsilon_{\phi}^{\rm nr}, which happens at χ=2\chi=2, or

Tmaxmν=(2π2SϕI3)1/3.\frac{T^{\max}}{m_{\nu}}=\left(\frac{2\pi^{2}}{S_{\phi}I_{3}}\right)^{1/3}.

Hence the maximum is ϵmax=mν/2\epsilon^{\max}=m_{\nu}/2. When SϕS_{\phi} increases (correspondingly mϕm_{\phi} decreases), the maximum ϵmax\epsilon^{\max} does not change significantly, but shifts to smaller TT.

Refer to caption
Figure 7: Dependence of comoving energy densities on T/mνT/m_{\nu} for different values of mϕ/(ymν)m_{\phi}/(ym_{\nu}). The dashed lines corresponds to the energy of the scalar field, the dotted lines to the energy of neutrinos, and the solid lines to the total energy.

In Fig. 7, we show the dependence of ϵϕ/mν\epsilon_{\phi}/m_{\nu} on TT (dashed lines). As can be easily checked, the analytic expressions above agrees well with the numerical results.

The energy density of neutrinos is given by

ρν(m~ν)=12π20𝑑pp2p2+m~ν2f(p,T).\rho_{\nu}(\tilde{m}_{\nu})=\frac{1}{2\pi^{2}}\int_{0}^{\infty}dp\leavevmode\nobreak\ p^{2}\sqrt{p^{2}+\tilde{m}_{\nu}^{2}}f(p,T).

It can be rewritten as

ρν(m~ν)=12π2[0dpp4p2+m~ν2f(p,T)+m~ν20dpp2p2+m~ν2f(p,T)].\rho_{\nu}(\tilde{m}_{\nu})=\frac{1}{2\pi^{2}}\left[\int_{0}^{\infty}\frac{dp\leavevmode\nobreak\ p^{4}}{\sqrt{p^{2}+\tilde{m}_{\nu}^{2}}}f(p,T)+\tilde{m}_{\nu}^{2}\int_{0}^{\infty}\frac{dp\leavevmode\nobreak\ p^{2}}{\sqrt{p^{2}+\tilde{m}_{\nu}^{2}}}f(p,T)\right]. (85)

Here the second integral coincides with integral in ν¯ν\langle\bar{\nu}\nu\rangle and ϕ\phi. The dependence of ρν(m~ν)\rho_{\nu}(\tilde{m}_{\nu}) on TT can be found explicitly.

In the non-relativistic case we obtain from Eq. (85)

ρνnr=12π2[m~νI3T3+1m~νI5T5]12π2mνI3T3,\rho_{\nu}^{\rm nr}=\frac{1}{2\pi^{2}}\left[\tilde{m}_{\nu}I_{3}T^{3}+\frac{1}{\tilde{m}_{\nu}}I_{5}T^{5}\right]\approx\frac{1}{2\pi^{2}}m_{\nu}I_{3}T^{3},

and consequently, ρνnr=mνnν\rho_{\nu}^{\rm nr}=m_{\nu}n_{\nu}, or

ϵνnr=mν.\epsilon_{\nu}^{\rm nr}=m_{\nu}.

In the relativistic case, we have

ρνrel=12π2[I4T4+m~ν2I2T2]I42π2T47π2240T4,\rho_{\nu}^{\rm rel}=\frac{1}{2\pi^{2}}\left[I_{4}T^{4}+\tilde{m}_{\nu}^{2}I_{2}T^{2}\right]\approx\frac{I_{4}}{2\pi^{2}}T^{4}\approx\frac{7\pi^{2}}{240}T^{4},

and

ϵνrel=I4I3T3.15T,\epsilon_{\nu}^{\rm rel}=\frac{I_{4}}{I_{3}}T\approx 3.15T,

as expected. The dependence of ϵν\epsilon_{\nu} on TT is shown in Fig. 7 (dotted lines).

Solid lines in Fig. 7 correspond to the total energy (80). Scalar contributions shift the minimum of ϵtot\epsilon_{\rm tot} to slightly larger TT with respect to the minimum of ϵν\epsilon_{\nu}.

The key feature of the ϵtot\epsilon_{\rm tot} dependence on TT is development of the dip at some temperature below mνm_{\nu}, denoted by TdipT_{\rm dip}.

According to Fig. 7, ϵtot\epsilon_{\rm tot} becomes smaller than mνm_{\nu} when Tmν/3T\simeq m_{\nu}/3. As TT decreases below TdipT_{\rm dip}, the energy ϵtot\epsilon_{\rm tot} converges to mνm_{\nu} from below. The dip essentially coincides with the transition region for m~ν\tilde{m}_{\nu} (Fig. 6).

Without the Yukawa interaction (or if the Yukawa interaction is too weak, as illustrated by the red line for Sϕ=0.042=625S_{\phi}=0.04^{-2}=625 in Fig. 7), ϵtot\epsilon_{\rm tot} would decrease as ϵtotT\epsilon_{\rm tot}\propto T in the relativistic case; at Tmν/3T\simeq m_{\nu}/3, the curve flattens and ϵtot\epsilon_{\rm tot} converges to mνm_{\nu}. For the entire range it would always be ϵtot>mν\epsilon_{\rm tot}>m_{\nu} which corresponds to unbounded neutrinos. For larger SϕS_{\phi} (other curves in Fig. 7), the dip appears. The appearance of the dip with ϵtot<mν\epsilon_{\rm tot}<m_{\nu} means that neutrinos can form bound systems with the average binding energy equal to mνϵtotm_{\nu}-\epsilon_{\rm tot}. As the interaction strength increases, the dip shifts to lower temperatures and becomes deeper. Correspondingly the binding energy becomes stronger.

The critical value of the strength needed for existence of the dip, SϕcS_{\phi}^{c}, is determined by the condition that temperature of beginning of deviation of ϵνrel\epsilon_{\nu}^{\rm rel} from linear decrease equals Trelmν/3T_{\rm rel}\approx m_{\nu}/3. This temperature coincides with TrelT_{\rm rel} of deviation of m~ν\tilde{m}_{\nu} from its 1/T21/T^{2} behavior. The condition (78) can be rewritten as

Sϕc24(mνTrel)3.S_{\phi}^{c}\approx 24\left(\frac{m_{\nu}}{T_{\rm rel}}\right)^{3}.

Then for Trel/mν1/3T_{\rm rel}/m_{\nu}\approx 1/3 it gives Sϕc=580S_{\phi}^{c}=580, which is close to Sϕ=625S_{\phi}=625 for the red curve in Fig. 7. In actual numerical calculations, we find that the dip appears roughly at Sϕc700S_{\phi}^{c}\approx 700.

VI.4 Fragmentation

Here we consider Sϕ>SϕcS_{\phi}>S_{\phi}^{c} so that the dip occurs. Above TdipT_{\rm dip}, cosmic neutrinos evolve as a uniform background with decreasing density and effective temperature. For T<TdipT<T_{\rm dip}, further cosmological expansion assuming the uniform neutrino background would imply that the energy of the system increases. Since there is no injection of energy into the ν\nu-ϕ\phi system from outside, it is energetically profitable that the uniform neutrino background fragments into parts (clusters) with the temperature TTdipT\approx T_{\rm dip}, which determines also the corresponding number density. Further evolution of fragments will proceed, subject to the dynamics of ν\nu clusters. The expansion of the Universe then increases the distance between fragments.

The highest temperature of fragmentation corresponds to Tf0.3mν0.03T_{f}\sim 0.3m_{\nu}\sim 0.03 eV. This temperature is achieved at redshift zf=200z_{f}=200 when the number density of neutrinos was

nf=9108cm3.n_{f}=9\cdot 10^{8}\,\,{\rm cm}^{-3}. (86)

In the process of formation of final configuration this density could further increase in the central parts of objects. This value agrees well with the one in Sec. VI.1.

The biggest possible object after fragmentation, which satisfies minimal (free) energy condition, would have the size D01/2DU(zf)D_{0}\leq 1/2D_{U}(z_{f}) in one dimension, where DU(zf)D_{U}(z_{f}) is the size of horizon in the epoch of fragmentation, DU(0)=4.2D_{U}(0)=4.2 Gpc. Therefore in three dimensions one would expect that 23=82^{3}=8 such objects could be formed, and for estimations we will use 10\sim 10 objects. At zf=200z_{f}=200 we find D0=10D_{0}=10 Mpc. In the present epoch distance between the objects is zfz_{f} times larger, that is 2000\sim 2000 Mpc.

Using the number density density (86) and the radius Rf=5R_{f}=5 Mpc we obtain the total number of neutrinos in a cluster

Nf=4π3Rf3nf=1.21085.N_{f}=\frac{4\pi}{3}R_{f}^{3}n_{f}=1.2\cdot 10^{85}. (87)

Correspondingly, the mass is M=41017MM=4\cdot 10^{17}M_{\odot}. These biggest structures can be realized for mϕ=41030eVm_{\phi}=4\cdot 10^{-30}{\rm eV} and y=41027y=4\cdot 10^{-27}.

If mϕm_{\phi} and yy are much larger, these structures do not satisfy conditions for stable bound systems. The number of neutrinos and radii should be much smaller. One can explore two possible ways of formation of small scale structures:

1. The neutrino background first fragments into the biggest structures, thus satisfying the energy requirement. These structures are unstable and further fragment into smaller parts. The secondary fragmentation can be triggered by density fluctuation of the neutrino background itself or via gravitational interactions with already existing structures like the Dark Matter halos.

If the secondary fragmentation is slow, the overall final structure could consist of superclusters (originated from primary fragmentation) and clusters within superclusters.

2. The uniform neutrino background may immediately fragment into small structures (close to final stable bound states) with R=𝒪(1/mϕ)R={\cal O}(1/m_{\phi}). These structures may interact among themselves.

As we established, for given mϕm_{\phi} and yy (i.e. fixed strength) there is a range of stable configurations with different NN and RR. E.g. for Sϕ1/2=0.01S_{\phi}^{-1/2}=0.01, NN can be within 4 orders of magnitude. What is the distribution of clusters with respect to NN? The answer depends on details of formation, in particular, on the surface effects. One can consider several options:

(i) a flat distribution within the allowed interval;

(ii) a peak at N=X(nF0)/y3N=X(n_{F0})/y^{3}, where nF0n_{F0} is the density at the beginning of fragmentation which, in turn, is determined by the strength, etc.

Considering TminT_{\min} as the temperature of the beginning of fragmentation we can find using Figs. 6 and 7 the physical conditions of medium which determine the initial states of pre-clusters. The kinetic energy per neutrino is (in the unit of mνm_{\nu}) ϵkin=3.14Tmin/mν\epsilon^{\rm kin}=3.14T_{\min}/m_{\nu}. The kinetic energy turns out to be much bigger than m~(Tmin)\tilde{m}(T_{\min}):

ϵkinm~(Tmin).\epsilon^{\rm kin}\gg\tilde{m}(T_{\min}).

So neutrinos are ultra-relativistic and therefore our usage of the relativistic distribution is justified. Correspondingly, the total energy in neutrinos nearly coincides with the kinetic energy: ϵνϵkin\epsilon_{\nu}\approx\epsilon^{\rm kin}.

Thus, for Sϕ1/2=103S_{\phi}^{-1/2}=10^{-3} we obtain from Figs. 6 and 7: Tmin/mν=5102T_{\min}/m_{\nu}=5\cdot 10^{-2}, ϵkin=0.15\epsilon^{\rm kin}=0.15, and m~/mν=0.01\tilde{m}/m_{\nu}=0.01. The energy in the scalar field is substantially smaller than in neutrinos: ϵϕ=0.05\epsilon_{\phi}=0.05. However it increases quickly with the temperature further decreasing. The total energy in the system is ϵtot=0.2\epsilon_{\rm tot}=0.2.

The number density of neutrinos is given by n0min=2106n_{0}^{\min}=2\cdot 10^{6} cm-3 which would correspond to the Fermi momentum pF/mν=0.1p_{F}/m_{\nu}=0.1. This momentum is smaller than the kinetic energy, so the cooling is needed to reach degenerate configuration.

Further results and statements should be considered as a possible trend since to construct the figures we used approximation of uniform infinite medium and relativistic distributions of neutrinos over momenta. Below TdipT_{\rm dip}, the dynamics of finite size objects should be included.

As TT deceases below TminT_{\min}, the kinetic energy of neutrinos decreases while the effective mass m~\tilde{m} increases. This means that the kinetic energy transforms into the increase of the mass as well as into the energy of scalar field. Thus, for T/mν=0.03T/m_{\nu}=0.03 we obtain ϵkin=0.09\epsilon^{\rm kin}=0.09, m~/mν=0.03\tilde{m}/m_{\nu}=0.03, i.e. the ratio ϵkinmν/m~=3\epsilon^{\rm kin}m_{\nu}/\tilde{m}=3; for T/mν=0.02T/m_{\nu}=0.02 the ratio becomes 0.3.

According to Figs. 6 and 7, the minimum of total neutrino energy, ϵν\epsilon_{\nu}, is achieved at a temperature lower than TdipT_{\rm dip}. In the minimum, for Sϕ1/2=103S_{\phi}^{-1/2}=10^{-3}, ϵν=0.1\epsilon_{\nu}=0.1 and m~/mν=0.07\tilde{m}/m_{\nu}=0.07, i.e. neutrinos become non-relativistic.

As TT further decreases, both ϵν\epsilon_{\nu} and m~\tilde{m} quickly increase with m~ϵν\tilde{m}\rightarrow\epsilon_{\nu}. This indicates that the kinetic energy is transformed into m~\tilde{m} as well as into ϵϕ\epsilon_{\phi}. One can guess that the phase transition ends when ϵϕ\epsilon_{\phi} reaches maximum which corresponds to the maximal potential energy. This happens at a slightly lower temperature than the temperature of ϵν\epsilon_{\nu} reaching its minimum.

The period of (primary) fragmentation can be estimated as the time interval which corresponds to the width of the dip:

Δt=t(T1)t(T2),\Delta t=t(T_{1})-t(T_{2}), (88)

where t23H(T)t\approx\frac{2}{3H(T)} (for matter dominated era) and HH is roughly proportional to T2T^{2}. The specific form can be determined by H2=8πρtotal/(3mpl2)H^{2}=8\pi\rho_{{\rm total}}/(3m_{{\rm pl}}^{2}) where ρtotal\rho_{{\rm total}} is the total energy density of the Universe and mplm_{{\rm pl}} is the Planck mass. The two temperatures T1T_{1} and T2T_{2} correspond to the maximum of the dashed curve in Fig. 7 and the minimum of the solid curve respectively. This gives e.g. Δt=0.014τU\Delta t=0.014\tau_{U} for Sϕ1/2=103S_{\phi}^{-1/2}=10^{-3} and Δt=0.24τU\Delta t=0.24\tau_{U} for Sϕ1/2=104S_{\phi}^{-1/2}=10^{-4}, where τU\tau_{U} is the present age of the Universe. Note that the estimations based on Fig. 7 may not be reliable since in the region of temperatures below TdipT_{\rm dip} neutrinos become non-relativistic and the dynamics of clusters starts to dominate.

According fo Fig. 7, TdipT_{\rm dip} decreases as SϕS_{\phi} increases. This means that for larger SϕS_{\phi}, fragmentation starts at later epochs and lower densities. For Sϕ1/2=(0.12, 102, 103)S_{\phi}^{-1/2}=(0.12,\leavevmode\nobreak\ 10^{-2},\leavevmode\nobreak\ 10^{-3}), we find Tmin/mν=(0.3, 0.12, 0.04)T_{\rm min}/m_{\nu}=(0.3,\leavevmode\nobreak\ 0.12,\leavevmode\nobreak\ 0.04), n0=(9108, 5.7107, 2.1106)n_{0}=(9\cdot 10^{8},\leavevmode\nobreak\ 5.7\cdot 10^{7},\leavevmode\nobreak\ 2.1\cdot 10^{6}) cm-3 and the redshift at fragmentation (zf+1)=(200, 80, 27)(z_{f}+1)=(200,\leavevmode\nobreak\ 80,\leavevmode\nobreak\ 27). With the increase of strength, n0n_{0} and pF0p_{F0} decrease. Then according to Figs. 5 and 4, NN and RR decrease.

Let us underline that results on the very fact of formation based on the energy consideration and properties of final configurations are robust. Details and features of fragmentation (the way of fragmentation, the distribution of clusters as a function of NN, the level of degeneracy in the final states of bound systems) require further studies, and in general, extensive numerical simulations.

VI.5 Virialization

In the range of Sϕ=70700S_{\phi}=70-700, the dip disappears and the phase transition is absent. Such interaction strengths still allow for stable bound systems of neutrinos. In this case, ν\nu clusters might be formed from the growth of local over-density to the non-linear regime and then go also through virialization, similar to the formation of dark matter halos.

Quantitatively, results for Yukawa forces (if mϕm_{\phi} is sufficiently small) can be obtained from the gravitational results by replacing

GM2r(yN)24πr,\frac{GM^{2}}{r}\rightarrow\frac{(yN)^{2}}{4\pi r}\thinspace, (89)

where GG is the gravitational constant and M=mνNM=m_{\nu}N is the total mass. The key difference from the gravitational case is that the strength of Yukawa interaction is many orders of magnitude larger than the strength of gravity.

With this analogy we can consider the following picture of structure formation.

1. The initial state is the neutrino background with some primordial or initial density perturbations (primordial clouds) with Δnp/np1\Delta n_{p}/n_{p}\ll 1 and sizes RpR_{p}. These perturbations may be related to the density perturbation of DM in the Universe, since neutrino structures are formed latter.

2. As the Universe expands, the size of clouds and distance between them increase simultaneously as the scale factor RdaR\propto d\propto a. Correspondingly, the density decreases as a3a^{-3} and Δn/n\Delta n/n is constant. This is the linear regime.

3. At certain epoch, when T<0.3mνT<0.3m_{\nu} the evolution becomes non-linear: the attraction becomes more efficient than expansion and the increase of the size of clouds becomes slower than the expansion of the Universe. This happens provided that the potential energy of the cloud is substantially larger than the total kinetic energy. Then the distance between two clouds increases faster that size of clouds. Also the decrease of the density slows down. 4. At some point (for which we denote the scale factor by amaxa_{\max}) the density stops decreasing, and after that (a>amaxa>a_{\max}), the system starts collapsing. In this way a bound system is formed and its evolution is largely determined by internal factors. The distance between two clouds continues increasing.

Since the system is collisionless and there is no energy loss (emission of particles or and waves) the contraction has a character of virialization777A system of collisionless (i.e. no additional interactions except for gravity) particles can collapse from a large cold distribution (or more strictly, a distribution with low mean kinetic energy since it is usually not a thermal distribution) to a smaller and hotter one. This processes is known as virialization., which starts from the stage that kinetic energy of the neutrino system becomes significantly smaller than the potential energy:

EK|EV|ξE1.\frac{E_{K}}{|E_{V}|}\equiv\xi_{E}\ll 1. (90)

During virialization, the kinetic energy increases (the effective TT increases, although the distribution may not be thermal). The process stops at a=avira=a_{\rm vir} when virial equilibrium is achieved, which also corresponds to hydrostatic equilibrium. The radius decreases by a factor of 2 so that the density increases by a factor of 8. Approaching the equilibrium the cluster may oscillate. At this point, the formation of the cluster is accomplished.

In what follows we will make some estimations and check how this picture matches the final configurations of clusters we obtained before. We will discuss bounds on initial parameters of the perturbations , formation procedure, epochs of different phases, etc, which follow from this matching.

The total energy is conserved during virialization. When the system reaches virial equilibrium, the virial theorem implies that

2EKvir+EVvir=0orEKvirEVvir=12,2E_{K}^{\rm vir}+E_{V}^{\rm vir}=0\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ {\rm or}\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \frac{E_{K}^{\rm vir}}{E_{V}^{\rm vir}}=\frac{1}{2}\thinspace, (91)

Here EKvirE_{K}^{\rm vir} and EVvirE_{V}^{\rm vir} are the total kinetic and potential energies after virialization. Thus, the ratio of energies increases.

Although quantitative studies of virialization involve NN-body simulation, the orders of magnitude of various characteristics can be obtained from simplified considerations. For gravitational virialization, it is known that888See, e.g., http://www.astro.yale.edu/vdbosch/astro610_lecture8.pdf or Ref. Mota and van de Bruck (2004). the virialization time (roughly the time of collapse) driven by gravity equals

tvir2πGM|M2(EK+EV)|3/2,t_{{\rm vir}}\simeq 2\pi GM\left|\frac{M}{2(E_{K}+E_{V})}\right|^{3/2}, (92)

where EKE_{K} and EVE_{V} are the total kinetic and potential energies (EK+EV<0E_{K}+E_{V}<0) of the system.

The time of mϕm_{\phi}-driven virialization can be obtained from Eq. (92) by making the replacement (89):

tvirmνN5/2y242|EK+EV|3/2.t_{{\rm vir}}\simeq\frac{\sqrt{m_{\nu}}N^{5/2}y^{2}}{4\sqrt{2}|E_{K}+E_{V}|^{3/2}}\thinspace. (93)

The contraction stops when virial equilibrium is achieved and we assume that the equilibrium has the same form as in the gravity case (91). Since the total energy is conserved during virialization,

EKvir+EVvir=EK+EV,E_{K}^{{\rm vir}}+E_{V}^{{\rm vir}}=E_{K}+E_{V}\thinspace,

we obtain EKvir=|EK+EV|E_{K}^{\rm vir}=|E_{K}+E_{V}| and EVvir=2|EK+EV|E_{V}^{\rm vir}=-2|E_{K}+E_{V}|.

The total energy in the beginning of virialization can be written as

|EK+EV|=|EV|(1ξE)=23|EV|,|E_{K}+E_{V}|=|E_{V}|(1-\xi_{E})=\frac{2}{3}|E_{V}|,

where in the last equality we have taken ξE=1/3\xi_{E}=1/3. Inserting this total energy with |EV||E_{V}| given in (52) into Eq. (93), we obtain the time of virialization

tvir3(52)3/2π2I31/21ymν(mνT)3/2,t_{{\rm vir}}\simeq\sqrt{3}\left(\frac{5}{2}\right)^{3/2}\frac{\pi^{2}}{I_{3}^{1/2}}\frac{1}{ym_{\nu}}\left(\frac{m_{\nu}}{T}\right)^{3/2},

which does not depend on NN. Numerically, we get

tvir3.4106sec(107y)(0.1eVmν)(mνT)3/2,t_{{\rm vir}}\simeq 3.4\cdot 10^{-6}{\rm sec}\left(\frac{10^{-7}}{y}\right)\left(\frac{0.1{\rm eV}}{m_{\nu}}\right)\left(\frac{m_{\nu}}{T}\right)^{3/2}, (94)

which implies that for T/mν=1T/m_{\nu}=1, the virialization takes tvir3.4106t_{{\rm vir}}\simeq 3.4\cdot 10^{-6} sec.

As aforementioned, after the virialization process, the radius decreases by a factor of 2 and correspondingly, the density increases by a factor of 8. Hence in the example we discussed at the end of the previous subsection, this gives Rvir=1.3R_{\rm vir}=1.3 km, and nvir=71010n_{\rm vir}=7\cdot 10^{10} cm-3.

As a result of virialization, the distribution reaches virial equilibrium with the virialized energies given by

EK(vir)=12EV(vir)=EKin+EVin.E_{K}^{({\rm vir})}=-\frac{1}{2}E_{V}^{({\rm vir})}=E_{K}^{\rm in}+E_{V}^{\rm in}. (95)

A halo of neutrinos is formed. In the absence of energy loss, its number density remains constant without being diluted by a3a^{-3} (aa is the scale factor) due to the cosmological expansion. That is, a halo of high-density cosmic neutrinos would be frozen till today.

Let us estimate temperatures of the ν\nu-clusters at different phases of formation. We can assume that deviation from the linear regime occurs at TmνT\sim m_{\nu}. Also we assume that before collapse the state of the system can be approximately described by thermal equilibrium and the border effects can be neglected. Therefore, we can use results obtained in sect. IV C. The expression for temperature (54) can be rewritten as

Tmν=5.3103(y3N)2/3ξE.\frac{T}{m_{\nu}}=5.3\cdot 10^{-3}(y^{3}N)^{2/3}\xi_{E}.

Let us take y3N=1y^{3}N=1 (see implications in the Fig. 4, 5) then T/mν=1T/m_{\nu}=1 (the scale factor aNLa_{NL}) corresponds to ξE=189\xi_{E}=189. For ξE=1/3\xi_{E}=1/3 we obtain T/mν=1.8103T/m_{\nu}=1.8\cdot 10^{-3}. For y3N=32y^{3}N=32, T/mν=1T/m_{\nu}=1 corresponds to ξE=19\xi_{E}=19 and ξE=1/3\xi_{E}=1/3 is achieved at T/mν=1.8102T/m_{\nu}=1.8\cdot 10^{-2}. If y3N=103y^{3}N=10^{3}, T/mν=1T/m_{\nu}=1 is realized at ξE=1.9\xi_{E}=1.9, and ξE=1/3\xi_{E}=1/3 at T/mν=0.18T/m_{\nu}=0.18.

VI.6 Neutrino structure of the Universe

The formation of the neutrino clusters leads to neutrino structure in the Universe with overdense areas and voids. For simplicity we consider that all the clouds have the same number of neutrinos NN.

In the present epoch the distance between neutrinos without clustering equals

d0=(nν0)1/3=0.2cm.d_{0}=(n_{\nu}^{0})^{1/3}=0.2\leavevmode\nobreak\ {\rm cm}.

If neutrinos cluster into clouds with NN neutrino in each, the distance between (centers of) these clusters is

d=d0N1/3.d=d_{0}N^{1/3}.

Taking, e.g., N=61022N=6\cdot 10^{22}, we obtain d=78d=78 km which is much larger than the radius of each cloud, R=0.62R=0.62 km given by Eq. (64).

The existence of such a structure could have an impact on the direct detection of relic neutrinos in future experiments such as the PTOLEMY Baracchini et al. (2018); Betti et al. (2019). The result also depends on the motion of the clusters with respect to the Earth. The motion of the Earth and the solar system will average the effect of these small structures. According to Eq. (45), in the non-relativistic regime the radius of cluster decreases with increase of NN as N1/3N^{-1/3}. Therefore the ratio of spacing (distance) and radius increases as N2/3\propto N^{2/3}:

dR1.1102d0mνy2N2/3.\frac{d}{R}\approx 1.1\cdot 10^{-2}\leavevmode\nobreak\ d_{0}m_{\nu}y^{2}N^{2/3}.

Taking the above example (d=78d=78 km, R=0.62R=0.62 km), we have d/R=126d/R=126. So, voids occupy most of the space. With decrease of yy, the radius of cluster increases and NN also increases (for fixed maximal density). As a result, d/R=d/R= const which is determined by the neutrino mass.

Let us estimate how small scale clusters may affect the direct detection of relic neutrinos assuming that clouds reached their final configurations. The distance needed for a detector to travel through to meet a ν\nu-cluster, in average, can be estimated as L=(σcnc)1L=(\sigma_{c}n_{c})^{-1}, where σcπR2\sigma_{c}\approx\pi R^{2} is the cross section of the cluster and nc=d3n_{c}=d^{-3} is the number density of clusters. Consequently,

L=dπ(dR)2.L=\frac{d}{\pi}\left(\frac{d}{R}\right)^{2}. (96)

Numerically, in our example, we obtain from (96) L=4105L=4\cdot 10^{5} km. Since the circumference of the Earth is 41044\cdot 10^{4} km, the Earth detector will cross the cluster every 10 days in average. Here we have not taken into account motion of clusters.

If the time of observation is much longer than 10 days, the total rate of events should be the same with and without clustering of neutrinos. This can be seen by noticing that the interaction rate when the detector is in the cloud is enhanced by a factor of nc/n0n_{c}/n_{0} while the rate of the detector being in a sphere of neutrinos is reduced by a factor of nc/n0n_{c}/n_{0}. Hence the time-averaged result will be the same as in the non-clustered case.

If the clusters are very big (e.g. comparable with the solar system or more) which is realized for very small coupling constant yy, the rate will be modified if the detector is in a void or in the overdense region of cluster.

VII Conclusions

Yukawa interactions of neutrinos with a new light scalar boson ϕ\phi are widely considered recently. One of the generic consequences of such interactions is the potential existence of stable bound states and bound systems of many neutrinos (ν\nu-clusters). Our study addresses these questions: whether and when do bound systems exist? what are the characteristic parameters of these systems and how are they determined? what are the observational (cosmological, laboratory, etc.) consequences of ν\nu-clusters?

We find that for two-neutrino bound states the Yukawa coupling yy and the mass of the scalar mϕm_{\phi} should satisfy the bound λy2mν/(8πmϕ)>0.84\lambda\equiv y^{2}m_{\nu}/(8\pi m_{\phi})>0.84. Eigenstates and eigenvalues of a 2ν2\nu bound system are obtained by solving the Schrödinger equation numerically. Analytic results (including the critical value of λ\lambda as well as the binding energy and the radius of the system) can be obtained approximately using the variation principle. The binding energy in a two-neutrino state is always small compared to the neutrino mass due to smallness of yy. For allowed values of couplings yy and masses of scalar mϕm_{\phi}, the bound state of two neutrinos would have the size greater than 101210^{12} cm. Bound states of sub-cm size are possible for 10 keV sterile neutrinos with coupling y>104y>10^{-4}. As long as the Yukawa coupling is below the perturbativity bound, neutrinos in a two-neutrino bound state are non-relativistic.

For NN-neutrino bound systems, the ground state corresponds to a system of degenerate Fermi gas in which the Yukawa attraction is balanced by the Fermi gas pressure. For non-relativistic neutrinos the density distribution in the ν\nu-cluster is described by the Lane-Emden equation. We re-derive this equation and solve it numerically. As a feature of the Lane-Emden equation, the solution always has a finite radius RR. The radius RR decreases as NN increases in the non-relativistic regime.

Unlike 2ν2\nu bound states, the NN-neutrino bound system can enter the relativistic regime with sufficiently large NN. We have derived a set of equations to describe the ν\nu-clusters in relativistic regime. In the non-relativistic limit they reduce to the L-E equation. In the relativistic regime, the Yukawa attraction acquires the mν/Em_{\nu}/E suppression which leads to a number of interesting consequences:

  • The effective number density of neutrinos, n~\tilde{n}, is suppressed. It deviates from the conventional number density nn significantly when the neutrino momentum is large;

  • The Fermi degenerate pressure is able to balance the attraction for any NN, thus preventing the collapse of the system in contrast to gravity;

  • The radius of system increases with NN, which implies that relativistic NN-body bound states cannot be much more compact than the non-relativistic ones.

  • As NN increases the radius of a ν\nu-cluster first decreases (in the non-relativistic regime) and then increases (in the relativistic regime), attaining a minimum when pF0p_{F0} is comparable to the effective neutrino mass.

  • There is a maximal central density 109\sim 10^{9} cm-3, which is determined by the neutrino mass.

  • For a given mϕm_{\phi} there is a minimal value of Ny3Ny^{3} for which stable configurations can be formed. This minimal value increases as the interaction strength Sϕy2mν2mϕ2S_{\phi}\equiv\frac{y^{2}m_{\nu}^{2}}{m_{\phi}^{2}} decreases.

  • For a given interaction strength, there is a minimal radius of the cluster.

  • The minimal interaction strength (maximal mϕm_{\phi} for fixed yy) required to bind neutrinos in a ν\nu-cluster is Sϕ70S_{\phi}\gtrsim 70.

We have investigated possible formation of the ν\nu-clusters from the uniform relic neutrino background as the Universe cools down.

  • For the interaction strength Sϕ700S_{\phi}\gtrsim 700, a dip (see Fig. 7) appears in the evolution of the total energy of the ν\nu-ϕ\phi system at some temperature below mνm_{\nu}. In this case, the formation of neutrino structures has a character of phase transition. The dip leads to the instability of the uniform ν\nu background, causing its fragmentation when TT further decreases. The typical size of final fragments is R(0.110)/mϕR\sim(0.1-10)/m_{\phi}. A detailed picture of fragmentation should be obtained by numerical simulations, which is beyond the scope of this work.

  • For 70Sϕ70070\lesssim S_{\phi}\lesssim 700, the dip is absent, and hence fragmentation does not occur, though stable neutrino clusters can exist. In this case neutrino clusters might be formed via the growth of local density perturbations followed by virialization, in a way similar to the formation of DM halos, though the energy loss of neutrino clusters due to ϕ\phi emission and neutrino annihilation is negligible. In this case, a new cooling mechanism would be required.

  • For Sϕ70S_{\phi}\lesssim 70, stable neutrino clusters cannot exist.

If clusters are formed at redshift zfz_{f} (which can be as large as 200) and retain the same size until today, the neutrino structure of the universe would show up as ν\nu-clusters which are zf3z_{f}^{3} times denser than the homogeneous neutrino background and voids which are zfz_{f} times larger than clusters. This leads to important implications for the detection of relic neutrinos.

While conclusion of formation of the neutrino clusters via fragmentation at Sϕ700S_{\phi}\gtrsim 700 is robust, details of fragmentation, the distribution of clusters with respect to NN, and the evolution of clusters require further studies.

Appendix A Calculation of ψ¯ψ\langle\overline{\psi}\psi\rangle and ψψ\langle\psi^{\dagger}\psi\rangle

The quantization of ψ\psi is formulated as follows:

ψ(x)=d3𝐩(2π)312E𝐩s[a𝐩sus(p)eipx+b𝐩svs(p)eipx].\psi(x)=\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{1}{\sqrt{2E_{\mathbf{p}}}}\sum_{s}\left[a_{\mathbf{p}}^{s}u^{s}(p)e^{-ip\cdot x}+b_{\mathbf{p}}^{s\dagger}v^{s}(p)e^{ip\cdot x}\right]. (97)

Now consider a single particle state which has been modulated by a wave-packet function w(𝐩)w(\mathbf{p}):

|w=d3𝐩(2π)3w(𝐩)2E𝐩|𝐩,s,|𝐩,s=2E𝐩a𝐩s|0,|w\rangle=\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{w(\mathbf{p})}{\sqrt{2E_{\mathbf{p}}}}|\mathbf{p},s\rangle,\ |\mathbf{p},s\rangle=\sqrt{2E_{\mathbf{p}}}a_{\mathbf{p}}^{s\dagger}|0\rangle, (98)

so that it is localized in both the coordinate space and the momentum space. More explicitly, for

Ψ(𝐱)d3𝐩(2π)3w(𝐩)2E𝐩us(p)ei𝐩𝐱,\Psi(\mathbf{x})\equiv\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{w(\mathbf{p})}{\sqrt{2E_{\mathbf{p}}}}u^{s}(p)e^{i\mathbf{p}\cdot\mathbf{x}}, (99)

the probability of this particle appearing at position 𝐱\mathbf{x} is 𝒫(𝐱)|Ψ(𝐱)|2{\cal P}(\mathbf{x})\equiv|\Psi(\mathbf{x})|^{2} and we have chosen w(𝐩)w(\mathbf{p}) in a way that

𝒫(𝐱){finite𝐱[𝐱0Δ𝐱,𝐱0+Δ𝐱]0otherwise,w(𝐩){finite𝐩[𝐩0Δ𝐩,𝐩0+Δ𝐩]0otherwise,{\cal P}(\mathbf{x})\approx\begin{cases}{\rm finite}&\text{$\mathbf{x}$}\in[\mathbf{x}_{0}-\Delta\mathbf{x},\ \mathbf{x}_{0}+\Delta\mathbf{x}]\\ 0&{\rm otherwise}\end{cases},\ w(\mathbf{p})\approx\begin{cases}{\rm finite}&\mathbf{p}\in[\mathbf{p}_{0}-\Delta\mathbf{p},\ \mathbf{p}_{0}+\Delta\mathbf{p}]\\ 0&{\rm otherwise}\end{cases}, (100)

which is always possible as long as Δ𝐱Δ𝐩1\Delta\mathbf{x}\Delta\mathbf{p}\gg 1. In this way, we can say that the particle is located approximately at 𝐱0\mathbf{x}_{0} with an approximate momentum 𝐩0\mathbf{p}_{0}.

Next, let us consider the normalization:

1=𝒫(𝐱)d3𝐱.1=\int{\cal P}(\mathbf{x})d^{3}\mathbf{x}. (101)

Substituting Eq. (99) in 𝒫(𝐱)|Ψ(𝐱)|2{\cal P}(\mathbf{x})\equiv|\Psi(\mathbf{x})|^{2} and then in Eq. (101), we have

1\displaystyle 1 =d3xd3𝐤(2π)3d3𝐩(2π)3w(𝐤)w(𝐩)2E𝐤2E𝐩us(k)us(p)ei(𝐩𝐤)𝐱\displaystyle=\int d^{3}x\frac{d^{3}\mathbf{\mathbf{k}}}{(2\pi)^{3}}\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{w^{*}(\mathbf{k})w(\mathbf{p})}{\sqrt{2E_{\mathbf{k}}}\sqrt{2E_{\mathbf{p}}}}u^{s\dagger}(k)u^{s}(p)e^{i(\mathbf{p}-\mathbf{k})\cdot\mathbf{x}}
=d3𝐤(2π)3d3𝐩(2π)3(2π)3δ3(𝐩𝐤)w(𝐤)w(𝐩)2E𝐤2E𝐩us(k)us(p)\displaystyle=\int\frac{d^{3}\mathbf{k}}{(2\pi)^{3}}\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}(2\pi)^{3}\delta^{3}(\mathbf{p}-\mathbf{k})\frac{w^{*}(\mathbf{k})w(\mathbf{p})}{\sqrt{2E_{\mathbf{k}}}\sqrt{2E_{\mathbf{p}}}}u^{s\dagger}(k)u^{s}(p)
=d3𝐩(2π)3|w(𝐩)|2,\displaystyle=\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}|w(\mathbf{p})|^{2}, (102)

where it is useful to note that

us(p)us(p)=2E𝐩,us(p)¯us(p)=2mψ.u^{s\dagger}(p)u^{s}(p)=2E_{\mathbf{p}},\ \overline{u^{s}(p)}u^{s}(p)=2m_{\psi}. (103)

With the above setup, we can evaluate the mean value of ψ¯ψ\overline{\psi}\psi and ψψ\psi^{\dagger}\psi in the presence of the single particle state |w|w\rangle. First, note that

a𝐩s|w\displaystyle a_{\mathbf{p}}^{s}|w\rangle =a𝐩sd3p(2π)3w(𝐩)a𝐩s|0\displaystyle=a_{\mathbf{p}}^{s}\int\frac{d^{3}p^{\prime}}{(2\pi)^{3}}w(\mathbf{p}^{\prime})a_{\mathbf{p}^{\prime}}^{s\dagger}|0\rangle
=d3𝐩w(𝐩)δ3(𝐩𝐩)|0\displaystyle=\int d^{3}\mathbf{p}^{\prime}\thinspace w(\mathbf{p}^{\prime})\delta^{3}(\mathbf{p}-\mathbf{p}^{\prime})|0\rangle
=w(𝐩)|0.\displaystyle=w(\mathbf{p})|0\rangle. (104)

Hence for w|ψψ|w\langle w|\psi^{\dagger}\psi|w\rangle, using Eq. (97) and Eq. (104), we have

w|ψ(x)ψ(x)|w\displaystyle\langle w|\psi^{\dagger}(x)\psi(x)|w\rangle =d3𝐤(2π)3d3𝐩(2π)312E𝐤2E𝐩w|a𝐤sus(k)eikxa𝐩sus(p)eipx|w\displaystyle=\int\frac{d^{3}\mathbf{k}}{(2\pi)^{3}}\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{1}{\sqrt{2E_{\mathbf{k}}}\sqrt{2E_{\mathbf{p}}}}\langle w|a_{\mathbf{k}}^{s\dagger}u^{s\dagger}(k)e^{ik\cdot x}a_{\mathbf{p}}^{s}u^{s}(p)e^{-ip\cdot x}|w\rangle
=d3𝐤(2π)3d3𝐩(2π)3us(k)us(p)2E𝐤2E𝐩ei(kp)x0|w(𝐤)w(𝐩)|0\displaystyle=\int\frac{d^{3}\mathbf{k}}{(2\pi)^{3}}\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{u^{s\dagger}(k)u^{s}(p)}{\sqrt{2E_{\mathbf{k}}}\sqrt{2E_{\mathbf{p}}}}e^{i(k-p)\cdot x}\langle 0|w(\mathbf{k})^{*}w(\mathbf{p})|0\rangle
=Ψ(x)Ψ(x),\displaystyle=\Psi^{\dagger}(x)\Psi(x), (105)

where the spacetime-generalized wavefunction, Ψ(x)\Psi(x), is defined similar to Eq. (99) except that ei𝐩𝐱e^{i\mathbf{p}\cdot\mathbf{x}} is replaced by eipxe^{-ip\cdot x}.

Similarly, for w|ψ¯ψ|w\langle w|\overline{\psi}\psi|w\rangle, we obtain

w|ψ¯(x)ψ(x)|w=Ψ¯(x)Ψ(x).\langle w|\overline{\psi}(x)\psi(x)|w\rangle=\overline{\Psi}(x)\Psi(x). (106)

As has been formulated in Eq. (100), the particle at t=0t=0 is localized in the region 𝐱[𝐱0Δ𝐱,𝐱0+Δ𝐱]\mathbf{x}\in[\mathbf{x}_{0}-\Delta\mathbf{x},\ \mathbf{x}_{0}+\Delta\mathbf{x}]. So it is useful to inspect the average value of w|ψ¯ψ|w\langle w|\overline{\psi}\psi|w\rangle and w|ψψ|w\langle w|\psi^{\dagger}\psi|w\rangle. At t=0t=0, we have

w|ψψ|wt=0,𝐱averagedlocally\displaystyle\langle w|\psi^{\dagger}\psi|w\rangle\xrightarrow{t=0,\ \mathbf{x}\ {\rm averaged\ locally}} 1Δx3𝐱0Δ𝐱𝐱0Δ𝐱d3𝐱w|ψψ|w\displaystyle\frac{1}{\Delta x^{3}}\int_{\mathbf{x}_{0}-\Delta\mathbf{x}}^{\mathbf{x}_{0}-\Delta\mathbf{x}}d^{3}\mathbf{x}\langle w|\psi^{\dagger}\psi|w\rangle
=\displaystyle= 1Δx3+d3𝐱Ψ(𝐱)Ψ(𝐱)\displaystyle\frac{1}{\Delta x^{3}}\int_{-\infty}^{+\infty}d^{3}\mathbf{x}\Psi^{\dagger}(\mathbf{x})\Psi(\mathbf{x})
=\displaystyle= 1Δx3d3𝐤(2π)3d3𝐩(2π)3us(k)us(p)2E𝐤2E𝐩w(𝐤)w(𝐩)\displaystyle\frac{1}{\Delta x^{3}}\int\frac{d^{3}\mathbf{k}}{(2\pi)^{3}}\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{u^{s\dagger}(k)u^{s}(p)}{\sqrt{2E_{\mathbf{k}}}\sqrt{2E_{\mathbf{p}}}}w(\mathbf{k})^{*}w(\mathbf{p})
×[+d3𝐱ei(𝐩𝐤)𝐱]\displaystyle\times\left[\int_{-\infty}^{+\infty}d^{3}\mathbf{x}e^{i(\mathbf{p}-\mathbf{k})\cdot\mathbf{x}}\right]
=\displaystyle= 1Δx3d3𝐩(2π)3us(p)us(p)2E𝐩w(𝐩)w(𝐩)\displaystyle\frac{1}{\Delta x^{3}}\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{u^{s\dagger}(p)u^{s}(p)}{2E_{\mathbf{p}}}w(\mathbf{p})^{*}w(\mathbf{p})
=\displaystyle= 1Δx3,\displaystyle\frac{1}{\Delta x^{3}}, (107)

which is exactly the particle number density of a single particle distributed in the small volume Δx3\Delta x^{3}.

If ψ\psi^{\dagger} is replaced with ψ¯\overline{\psi} in Eq. (107), at the last step we would have us(p)¯us(p)2mψ\overline{u^{s}(p)}u^{s}(p)\rightarrow 2m_{\psi} [see Eq. (103)] and hence

w|ψψ|wt=0,𝐱averagedlocally1Δx3d3𝐩(2π)3mψE𝐩|w(𝐩)|2.\langle w|\psi^{\dagger}\psi|w\rangle\xrightarrow{t=0,\ \mathbf{x}\ {\rm averaged\ locally}}\frac{1}{\Delta x^{3}}\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{m_{\psi}}{E_{\mathbf{p}}}|w(\mathbf{p})|^{2}. (108)

Since |w(𝐩)|2|w(\mathbf{p})|^{2} is mostly concentrated near 𝐩0\mathbf{p}_{0} [see Eq. (100)], it can be approximately reduced to 1Δx3mψE𝐩0\frac{1}{\Delta x^{3}}\frac{m_{\psi}}{E_{\mathbf{p}_{0}}}, which in the non-relativistic regime is equivalent to the number density but in the relativistic regime is suppressed.

The above calculation can be straightforwardly generalized to NN particles with different momentum and difference spins. Each of them has an independent wave-packet function wi(𝐩)w_{i}(\mathbf{p}). In this case, we sum over the contributions together:

ψ¯ψ=i=1Nwi|ψ¯ψ|wi,ψψ=i=1Nwi|ψψ|wi.\langle\overline{\psi}\psi\rangle=\sum_{i=1}^{N}\langle w_{i}|\overline{\psi}\psi|w_{i}\rangle,\ \ \langle\psi^{\dagger}\psi\rangle=\sum_{i=1}^{N}\langle w_{i}|\psi^{\dagger}\psi|w_{i}\rangle. (109)

Denote

f(𝐩)=i=1N1Δx3|wi(𝐩)|2,f(\mathbf{p})=\sum_{i=1}^{N}\frac{1}{\Delta x^{3}}|w_{i}(\mathbf{p})|^{2}, (110)

which can be regarded as the momentum distribution function of the NN particles. Then the summation gives

ψψ\displaystyle\langle\psi^{\dagger}\psi\rangle =d3𝐩(2π)3f(𝐩),\displaystyle=\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}f(\mathbf{p}), (111)
ψ¯ψ\displaystyle\langle\overline{\psi}\psi\rangle =d3𝐩(2π)3mψE𝐩f(𝐩).\displaystyle=\int\frac{d^{3}\mathbf{p}}{(2\pi)^{3}}\frac{m_{\psi}}{E_{\mathbf{p}}}f(\mathbf{p}). (112)

Appendix B Potential issues on the binding energy and the field energy

There is a potential issue regarding the binding energy of two particles in Sec. II: for two particles, the binding energy does not contain a factor of two. To address this issue, here we discuss the two-body problem from both the classical and modern (field theory) perspectives.

In the classical picture, we assume that each particle can be infinitely split to smaller particles. Denote the two particles as AA and BB. If we could remove an infinitely small percentage (ϵ\epsilon, ϵ1\epsilon\ll 1) of particle BB to r=r=\infty, the energy cost would be y24π1rermϕϵ\frac{y^{2}}{4\pi}\frac{1}{r}e^{-rm_{\phi}}\epsilon (due to the attraction of AA) plus the energy cost to overcome the attraction of BB. The latter part will be canceled out when all the small fractions of BB are re-assembled at r=r=\infty. So the total energy cost (i.e. binding energy) to separate AA and BB is

Ebinding=y24π1rermϕϵ=y24π1rermϕ.E_{{\rm binding}}=\frac{y^{2}}{4\pi}\frac{1}{r}e^{-rm_{\phi}}\sum\epsilon=\frac{y^{2}}{4\pi}\frac{1}{r}e^{-rm_{\phi}}. (113)

Hence we conclude that the total binding energy for a two-body system should take N=1N=1 in Eq. (14).

In addition to the classical picture, from the perspective of field theory, an equivalent interpretation is that the binding energy is stored in the field ϕ\phi. According to Eq. (6), the Hamiltonian for static ϕ\phi is

H=[12(ϕ)2+12mϕ2ϕ2]d3𝐱.H=\int\left[\frac{1}{2}(\nabla\phi)^{2}+\frac{1}{2}m_{\phi}^{2}\phi^{2}\right]d^{3}\mathbf{x}. (114)

For the case of two particles, we have

n(𝐱)\displaystyle n(\mathbf{x}) =δ3(𝐱𝐱A)+δ3(𝐱𝐱B),\displaystyle=\delta^{3}(\mathbf{x}-\mathbf{x}_{A})+\delta^{3}(\mathbf{x}-\mathbf{x}_{B}), (115)
ϕ(𝐱)\displaystyle\phi(\mathbf{x}) =y4π(1rAerAmϕ+1rBerBmϕ),\displaystyle=-\frac{y}{4\pi}\left(\frac{1}{r_{A}}e^{-r_{A}m_{\phi}}+\frac{1}{r_{B}}e^{-r_{B}m_{\phi}}\right), (116)

where 𝐱A\mathbf{x}_{A} and 𝐱B\mathbf{x}_{B} are coordinates of AA and BB, and rA,B=|𝐱|2|𝐱A,B|2r_{A,B}=\sqrt{|\mathbf{x}|^{2}-|\mathbf{x}_{A,B}|^{2}}. Although HH computed in this way is divergent, by comparing two cases with |𝐱A𝐱B|=r|\mathbf{x}_{A}-\mathbf{x}_{B}|=r and |𝐱A𝐱B|=|\mathbf{x}_{A}-\mathbf{x}_{B}|=\infty, the difference of HH is finite. Plugging Eq. (116) into Eq. (114) and using (ϕ)2d3𝐱=ϕ2ϕd3𝐱\int(\nabla\phi)^{2}d^{3}\mathbf{x}=-\int\phi\nabla^{2}\phi d^{3}\mathbf{x}, it is straightforward to obtain the energy difference:

V=ΔH=y24π1rermϕ.V=\Delta H=-\frac{y^{2}}{4\pi}\frac{1}{r}e^{-rm_{\phi}}. (117)

This is consistent with Eq. (113) in the classical picture.

Appendix C Numerical method to solve the two-particle Schrödinger equation

For a quantum system containing two interacting particles, the Schrödinger equation reads

itΨ(𝐫1,𝐫2)=HΨ(𝐫1,𝐫2),i\frac{\partial}{\partial t}\Psi(\mathbf{r}_{1},\ \mathbf{r}_{2})=H\Psi(\mathbf{r}_{1},\ \mathbf{r}_{2}), (118)

where Ψ(𝐫1,𝐫2)\Psi(\mathbf{r}_{1},\ \mathbf{r}_{2}) is the wave function with 𝐫1\mathbf{r}_{1} and 𝐫2\mathbf{r}_{2} being the coordinates of the two particles, and HH is the Hamiltonian:

H=12m11212m222+V(𝐫1,𝐫2).H=-\frac{1}{2m_{1}}\nabla_{1}^{2}-\frac{1}{2m_{2}}\nabla_{2}^{2}+V(\mathbf{r}_{1},\ \mathbf{r}_{2}). (119)

Here m1m_{1} and m2m_{2} are the masses of the two particles and VV is the potential.

Define the reduced mass of the two particles:

μm1m2m1+m2,\mu\equiv\frac{m_{1}m_{2}}{m_{1}+m_{2}}, (120)

and

𝐫𝐫1𝐫2,𝐑m1𝐫1+m2𝐫2m1+m2,\mathbf{r}\equiv\mathbf{r}_{1}-\mathbf{r}_{2},\ \mathbf{R}\equiv\frac{m_{1}\mathbf{r}_{1}+m_{2}\mathbf{r}_{2}}{m_{1}+m_{2}}, (121)

where 𝐑\mathbf{R} is the coordinate of the center of mass. Then 1\nabla_{1} and 2\nabla_{2} can be replaced by the derivatives with respect to 𝐫\mathbf{r} and 𝐑\mathbf{R} (denoted as R\nabla_{R} and r\nabla_{r} below):

1=μm2R+r,2=μm1Rr,\nabla_{1}=\frac{\mu}{m_{2}}\nabla_{R}+\nabla_{r},\ \nabla_{2}=\frac{\mu}{m_{1}}\nabla_{R}-\nabla_{r}, (122)

So the Hamiltonian can be transformed to

H=HR+Hr,H=H_{R}+H_{r}, (123)

where

HR12(m1+m2)R2,Hr12μr2+V(r).H_{R}\equiv-\frac{1}{2\left(m_{1}+m_{2}\right)}\nabla_{R}^{2},\ \ H_{r}\equiv-\frac{1}{2\mu}\nabla_{r}^{2}+V(r). (124)

Note that here we assume that VV depends on rr only. Eqs. (123) and (124) allows us to separate the motion of 𝐑\mathbf{R} with that of 𝐫\mathbf{r}:

Ψ(𝐫1,𝐫2)=ψ(𝐫)ψ~(𝐑).\Psi(\mathbf{r}_{1},\ \mathbf{r}_{2})=\psi(\mathbf{r})\tilde{\psi}(\mathbf{R}). (125)

In this work we are only concerned with the part of ψ(𝐫)\psi(\mathbf{r}), which can be expanded using spherical harmonics. Plugging Eq. (21) to

Hrψ(𝐫)=Eψ(𝐫),H_{r}\psi(\mathbf{r})=E\psi(\mathbf{r}), (126)

we obtain

12μd2u(r)dr2+[V(r)+12μl(l+1)r2]u(r)=Eu(r).-\frac{1}{2\mu}\frac{d^{2}u(r)}{dr^{2}}+\left[V(r)+\frac{1}{2\mu}\frac{l(l+1)}{r^{2}}\right]u(r)=Eu(r). (127)

Define

V~2μV(r)+l(l+1)r2,E~2μE.\tilde{V}\equiv 2\mu V(r)+\frac{l(l+1)}{r^{2}},\ \tilde{E}\equiv 2\mu E. (128)

Then the equation is simplified to

u′′(r)+V~(r)u(r)=E~u(r).-u^{\prime\prime}(r)+\tilde{V}(r)u(r)=\tilde{E}u(r). (129)

Given any form of V~(r)\tilde{V}(r), in principle Eq. (129) can be numerically solved. Since rr varies from zero to infinity, it is actually more convenient to make the following variable transformation:

v(ω)=u(r)cosω,ωarctanrR,v(\omega)=u(r)\cos\omega,\ \omega\equiv\arctan\frac{r}{R}, (130)

so that v(ω)v(\omega) is a function on a finite interval ω[0,π/2)\omega\in[0,\pi/2). Here RR in principle can be an arbitrary length scale but in order to improve the efficiency of the numerical method it should be set at the typical length scale of the wave function that is being inspected.

After the transformation, Eq. (129) becomes

cos4ωR2d2dω2v(ω)+[V~(ω)cos4ωR2]v(ω)=E~v(ω).-\frac{\cos^{4}\omega}{R^{2}}\frac{d^{2}}{d\omega^{2}}v(\omega)+\left[\tilde{V}(\omega)-\frac{\cos^{4}\omega}{R^{2}}\right]v(\omega)=\tilde{E}v(\omega). (131)

This equation can be solved by identifying it as the problem of find eigenvalues and eigenvectors of a large matrix. If we divide the interval [0,π/2)[0,\pi/2) into NN segments and ω\omega takes one of these values:

(0,dω, 2dω,,π2dω,π2),\left(0,\ d\omega,\ 2d\omega,\ \cdots,\ \frac{\pi}{2}-d\omega,\ \frac{\pi}{2}\right), (132)

where dω=π2Nd\omega=\frac{\pi}{2N}, then d2dω2\frac{d^{2}}{d\omega^{2}} can be regarded as an operator in this discrete space, represented by the (N+1)×(N+1)(N+1)\times(N+1) matrix below:

d2dω21(π2N)2[2112112112112](N+1)×(N+1).\frac{d^{2}}{d\omega^{2}}\rightarrow\frac{1}{(\frac{\pi}{2N})^{2}}\left[\begin{array}[]{cccccc}-2&1\\ 1&-2&1\\ &1&-2&1\\ &&&&\cdots\\ &&&1&-2&1\\ &&&&1&-2\end{array}\right]_{(N+1)\times(N+1)}. (133)

And V~(ω)\tilde{V}(\omega) and cos4ω\cos^{4}\omega are represented by diagonal matrices:

V~(ω)diag[V~(0),V~(dω),V~(2dω),,V~(π2)],\tilde{V}(\omega)\rightarrow{\rm diag}\left[\tilde{V}(0),\ \tilde{V}(d\omega),\ \tilde{V}(2d\omega),\ \cdots,\ \tilde{V}(\frac{\pi}{2})\right], (134)
cos4ωdiag[cos4(0),cos4(dω),cos4(2dω),,cos4(π2)].\cos^{4}\omega\rightarrow{\rm diag}\left[\cos^{4}(0),\ \cos^{4}(d\omega),\ \cos^{4}(2d\omega),\ \cdots,\ \cos^{4}(\frac{\pi}{2})\right]. (135)

In the large-NN limit, by solving eigenvalues and eigenvectors of H~\tilde{H} defined below

H~cos4ωR2d2dω2+[V~(ω)cos4ωR2],\tilde{H}\equiv-\frac{\cos^{4}\omega}{R^{2}}\frac{d^{2}}{d\omega^{2}}+\left[\tilde{V}(\omega)-\frac{\cos^{4}\omega}{R^{2}}\right], (136)

we can obtain energy levels and wave functions.

Appendix D Known constraints on the yukawa coupling

There have been a variety of experimental bounds on a light scalar coupled to neutrinos. Here we compile known results (as summarized in Tab. 2) in the literature and comment on their relevance to the scenario considered in this work.

Table 2: A summary of known bounds on the neutrinophilic yukawa coupling yy, assuming mϕm_{\phi} is well below the neutrino mass.
processes flavor dependence bounds Ref.xx
π±\pi^{\pm} decay νe\nu_{e} y<1.3×102y<1.3\times 10^{-2} Berryman et al. (2018)
K±K^{\pm} decay νe\nu_{e}, νμ\nu_{\mu} y<1.4×102y<1.4\times 10^{-2} (νe\nu_{e}) or <3×103<3\times 10^{-3} (νμ\nu_{\mu}) Berryman et al. (2018)
ββ\beta\beta decay νe\nu_{e} y<3.4×105y<3.4\times 10^{-5} Agostini et al. (2015)
ZZ decay all flavors y<0.3y<0.3 Brdar et al. (2020)
BBN all flavors y<4.6×106y<4.6\times 10^{-6} Huang et al. (2018)
CMB all flavors y<8.2×107y<8.2\times 10^{-7} Forastieri et al. (2015)
SN1987A (energy loss)xxxx all flavors y<3×107y<3\times 10^{-7} or 2×105<y<3×1042\times 10^{-5}<y<3\times 10^{-4} Kachelriess et al. (2000)
SN1987A (deleptonization) νe\nu_{e} y<2×106y<2\times 10^{-6} Farzan (2003)

In laboratory experiments, a light scalar that exclusively couples to neutrinos can be constrained by particle physics processes such as meson decay (π±\pi^{\pm}, K±K^{\pm}Barger et al. (1982); Lessa and Peres (2007); Pasquini and Peres (2016); Berryman et al. (2018), ZZ invisible decay Brdar et al. (2020), and double beta (ββ\beta\beta) decay Burgess and Cline (1993, 1994); Blum et al. (2018); Cepedello et al. (2019); Brune and Paes (2019); Deppisch et al. (2020). Since the typical momentum transfer scales of these processes varying from 10210^{2} GeV to MeV are well above the mass of ϕ\phi considered in this work, the constraints from these processes on the coupling yy are almost independent of mϕm_{\phi} when mϕm_{\phi} is lighter than mνm_{\nu}. Most of these particle decay bounds only apply to the electron and/or muon flavor, except for ZZ invisible decay which is relevant to all flavors (For couplings to ντ\nu_{\tau}, τ\tau decay is less restrictive than ZZ decay Brdar et al. (2020)). In addition, the ββ\beta\beta decay bound is only valid when ϕ\phi is coupled to two electron neutrinos instead of one neutrino and one anti-neutrino. The specific values of aforementioned bounds are summarized in Tab. 2. High precision measurements of neutrino scattering in principle could be sensitive to such interactions via bremsstrahlung or loop processes, which compared to the standard processes are typically suppressed by a factor of y2/16π2102y2y^{2}/16\pi^{2}\sim 10^{-2}y^{2}. Current neutrino scattering experiments can only measure the cross sections at percent-level precision Bilmis et al. (2015); Lindner et al. (2018), corresponding to an upper limit of y1y\lesssim 1, which is negligibly weak. Long-range force searches (also referred to as the fifth force searches in the literature) based on precision test of gravitational laws (e.g. the inverse-square law, weak equivalent principle) are sensitive to light mediators coupled to normal matter (electrons or baryons) Adelberger et al. (2007, 2009); Wagner et al. (2012). Although in the presence of ϕ\phi-neutrino couplings, ϕ\phi-electron couplings can be induced at the one-loop level Xu (2020); Chauhan and Xu (2021), such loop-induced couplings are typically highly suppressed by the large hierarchy between neutrino masses and the electroweak scale. Hence bounds from long-range force searches can be neglected in this work.

Astrophysical and cosmological bounds, by contrast, are much more restrictive. Observations of neutrinos from the supernova event SN1987A have been employed in the literature to set various constraints from different considerations. The most commonly considered effect is energy loss, which leads to the constraint y<3×107y<3\times 10^{-7} or 2×105<y<3×1042\times 10^{-5}<y<3\times 10^{-4} Kachelriess et al. (2000). If the new interaction is able to convert νe\nu_{e} to ν¯e\overline{\nu}_{e} or to neutrinos of other flavors, then another effect known as deleptonization can also put a strong bound, y<2×106y<2\times 10^{-6} Farzan (2003). Other supernova bounds Dent et al. (2012); Kazanas et al. (2014); Heurtier and Zhang (2017); Dev et al. (2020) assuming the presence of quark couplings, heavy masses, or electromagnetic decays do not apply here. In addition, for y>4.6×106y>4.6\times 10^{-6} the scalar could be thermalized in the early universe, causing observable effects in the Big Bang Nucleosynthesis (BBN) Huang et al. (2018). For smaller yy, the scalar cannot be thermalized but could affect the evolution of cosmological perturbations and hence be constrained by CMB data Forastieri et al. (2015). Theses bounds are also summarized in Tab. 2.

Appendix E Equivalence between chemical equilibrium and force balance

In this appendix, we show that the balance of the two forces, FYuk=FdegF_{{\rm Yuk}}=F_{{\rm deg}} in Eq. (39), is equivalent to chemical equilibrium of the system (equilibrium of particle numbers in the presence of external forces). Indeed, the chemical equilibrium of system means that

dμdr=0,\frac{d\mu}{dr}=0, (137)

where μ\mu is the chemical potential in the Fermi-Dirac distribution f=[e(Eμ)/T+1]1f=[e^{(E-\mu)/T}+1]^{-1} with the neutrino energy EE modified by ϕ\phi:

E=(mν+yϕ)2+p2.E=\sqrt{(m_{\nu}+y\phi)^{2}+p^{2}}. (138)

The Fermi energy, EFE_{F}, is defined as Eq. (138) with p=pFp=p_{F}. Since it is the maximal energy of a particle in the distribution, EF=μE_{F}=\mu, Eq. (137) is equivalent to dEF/dr=0dE_{F}/dr=0. So we have

ddr[(mν+yϕ)2+pF2]=0,\frac{d}{dr}\left[(m_{\nu}+y\phi)^{2}+p_{F}^{2}\right]=0, (139)

or

y(mν+yϕ)dϕdr+pFdpFdr=0.y(m_{\nu}+y\phi)\frac{d\phi}{dr}+p_{F}\frac{dp_{F}}{dr}=0. (140)

Let us show that in nonrelativistic limit dϕdr\frac{d\phi}{dr} and dpFdr\frac{dp_{F}}{dr} can be related to the Yukawa force and the degenerate pressure respectively.

To make use of the classic concept of forces, we need to take the non-relativistic limit in which yϕmνy\phi\ll m_{\nu} so that yϕy\phi can be neglected in the first term of (140). The gradient dϕdr\frac{d\phi}{dr} is related to the Yukawa force as

FYuk(r)yndϕdr.F_{{\rm Yuk}}(r)\equiv-yn\frac{d\phi}{dr}. (141)

Therefore from Eq. (140) we have

FYuk(r)=npFmνdpFdr.F_{{\rm Yuk}}(r)=n\frac{p_{F}}{m_{\nu}}\frac{dp_{F}}{dr}. (142)

According to Eq. (40), it is straightforward to see that in the non-relativistic case

d𝒫degdr=npFmνdpFdr.\frac{d{\cal P}_{{\rm deg}}}{dr}=n\frac{p_{F}}{m_{\nu}}\frac{dp_{F}}{dr}. (143)

Thus, in the non-relativistic regime, Eq. (137) and Eq. (39) are equivalent to each other.

Appendix F Derivation of the Lane-Emden equation from equations of motion

For non-relativistic neutrinos without interactions, we have E=mν2+p2mν+p22mνE=\sqrt{m_{\nu}^{2}+p^{2}}\approx m_{\nu}+\frac{p^{2}}{2m_{\nu}} and hence Eμ0E-\mu\leq 0 corresponds to mν+p22mνμ0m_{\nu}+\frac{p^{2}}{2m_{\nu}}-\mu\lesssim 0. When p=pFp=p_{F}, Eμ0E-\mu\leq 0 is saturated. In the presence of the ϕ\phi-induced potential VmνV\ll m_{\nu}, according to Eq. (9), Eμ0E-\mu\leq 0 corresponds to mν+p22mν+Vμ0m_{\nu}+\frac{p^{2}}{2m_{\nu}}+V-\mu\lesssim 0. Hence it is convenient to define an effective chemical potential to absorb mνm_{\nu} and VV:

μeffmνV+μ.\mu_{{\rm eff}}\equiv-m_{\nu}-V+\mu\thinspace. (144)

When μeff\mu_{{\rm eff}} is positive, we have pFp_{F} determined by

pF22mνμeff0.\frac{p_{F}^{2}}{2m_{\nu}}-\mu_{{\rm eff}}\approx 0\thinspace. (145)

Note that μeff\mu_{{\rm eff}} sometimes can be negative, for which we should take pF=0p_{F}=0:

pF={2mνμeffif μeff>00if μeff<0.p_{F}=\begin{cases}\sqrt{2m_{\nu}\mu_{{\rm eff}}}&\text{if }\mu_{{\rm eff}}>0\\ 0&\text{if }\mu_{{\rm eff}}<0\end{cases}\thinspace. (146)

Consider Eq. (12) where the r.h.s depends on the local number density n(𝐱)n(\mathbf{x}), which according to Eq. (31) is determined by the local Fermi momentum pF(𝐱)p_{F}(\mathbf{x}), which according to Eq. (145) or (146) depends on VV. So eventually the r.h.s of Eq. (12), which as a source term generates the potential VV, is determined by VV itself:

[2mϕ2]V=y26π2[2mν(μmνV)]3/2.\left[\nabla^{2}-m_{\phi}^{2}\right]V=\frac{y^{2}}{6\pi^{2}}\left[2m_{\nu}(\mu-m_{\nu}-V)\right]^{3/2}. (147)

It is more convenient to express Eq. (147) in terms of μeff\mu_{{\rm eff}} using Eq. (144):

2μeff+mϕ2(μμeffmν)=y2(2mν)3/26π2μeff3/2.\nabla^{2}\mu_{{\rm eff}}+m_{\phi}^{2}(\mu-\mu_{{\rm eff}}-m_{\nu})=-\frac{y^{2}(2m_{\nu})^{3/2}}{6\pi^{2}}\mu_{{\rm eff}}^{3/2}. (148)

For spherically symmetric distributions and mϕ=0m_{\phi}=0, it becomes

1r2ddr[r2dμeff(r)dr]=κ~[μeff(r)]γ,\frac{1}{r^{2}}\frac{d}{dr}\left[r^{2}\frac{d\mu_{{\rm eff}}(r)}{dr}\right]=-\tilde{\kappa}\thinspace\left[\mu_{{\rm eff}}(r)\right]^{\gamma}, (149)

where

κ~y2(2mν)3/26π2,γ=32.\tilde{\kappa}\equiv\frac{y^{2}(2m_{\nu})^{3/2}}{6\pi^{2}},\ \ \gamma=\frac{3}{2}. (150)

When μeff\mu_{{\rm eff}} is expressed in terms of nn, one obtains Eq. (42).

Appendix G Numerical details of solving Eqs. (60) and (61)

It is useful to note that we can remove the dependence of the differential equation on yy by defining

¯/y,r¯yr,m¯ϕmϕ/y,\overline{\nabla}\equiv\nabla/y,\ \ \overline{r}\equiv yr,\ \ \overline{m}_{\phi}\equiv m_{\phi}/y, (151)

and rewrite Eq. (60) as

¯2m~m¯ϕ2(m~mν)\displaystyle\overline{\nabla}^{2}\tilde{m}-\overline{m}_{\phi}^{2}(\tilde{m}-m_{\nu}) =n~,\displaystyle=\tilde{n}, (152)

where

¯2m~=[d2dr¯2+2ddr¯]m~.\overline{\nabla}^{2}\tilde{m}=\left[\frac{d^{2}}{d\overline{r}^{2}}+2\frac{d}{d\overline{r}}\right]\tilde{m}. (153)

Eq. (61) is equivalent to

m~2(r¯)+pF2(r¯)=m~R2,\tilde{m}^{2}(\overline{r})+p_{F}^{2}(\overline{r})=\tilde{m}_{R}^{2}\thinspace, (154)

where m~R\tilde{m}_{R} is m~ν\tilde{m}_{\nu} at r=Rr=R (or r¯=R¯Ry\overline{r}=\overline{R}\equiv Ry). This allows us to substitute m~ν(r¯)m~R2pF2(r¯)\tilde{m}_{\nu}(\overline{r})\rightarrow\sqrt{\tilde{m}_{R}^{2}-p_{F}^{2}(\overline{r})} in Eqs. (152) and express it as an equation of pF(r¯)p_{F}(\overline{r}).

The initial values in Eq. (62) are determined as follows. We note that m~0\tilde{m}_{0} cannot be set arbitrarily because at rr\to\infty it should match mνm_{\nu}. Technically, for an arbitrarily m~0\tilde{m}_{0}, quite often one gets a solutions with pF(r)p_{F}(r) monotonically increasing or oscillating as rr increases, which is certainly not a physical solution. We find that in practice it is more convenient to use m~R\tilde{m}_{R} instead of m~0\tilde{m}_{0}. Hence, in our code implementation, we first set values of (m~R,pF0)(\tilde{m}_{R},\ p_{F0}) and then use m~0=m~R2pF02\tilde{m}_{0}=\sqrt{\tilde{m}_{R}^{2}-p_{F0}^{2}} to obtain (m~0,pF0)(\tilde{m}_{0},\ p_{F0}). After solving the differential equations for 0<r¯<R¯0<\overline{r}<\overline{R}, one can obtain N¯Ny3\overline{N}\equiv Ny^{3} and m~ν(r)\tilde{m}_{\nu}(r\to\infty) using to the following integrals:

N¯\displaystyle\overline{N} =0R¯n(r¯)4πr¯2𝑑r¯,\displaystyle=\int_{0}^{\overline{R}}n(\overline{r})4\pi\overline{r}^{2}d\overline{r}\thinspace, (155)
m~ν(r)\displaystyle\tilde{m}_{\nu}(r\to\infty) =m~R+em¯ϕR¯0R¯r¯n~(r¯)sinh(m¯ϕr¯)m¯ϕR¯𝑑r¯.\displaystyle=\tilde{m}_{R}+e^{-\overline{m}_{\phi}\overline{R}}\int_{0}^{\overline{R}}\overline{r}\tilde{n}(\overline{r})\frac{\sinh(\overline{m}_{\phi}\overline{r})}{\overline{m}_{\phi}\overline{R}}d\overline{r}\thinspace. (156)

With the above setup, we implement the code in the following way:

  • For each pair of the input parameter (m~R,pF0)(\tilde{m}_{R},\ p_{F0}), the code obtains a curve of pF(r¯)p_{F}(\overline{r}) which may or may not drop to zero.

  • By scanning over a range of m~R\tilde{m}_{R}, it can find the interval in which the former happens. In this interval, each solution has a radius R¯\overline{R} determined by the zero point of pF(r¯)p_{F}(\overline{r}).

  • Correspondingly, m~ν(r)\tilde{m}_{\nu}(r\to\infty) can be computed using Eq. (156). And one obtains a curve of m~ν(r)\tilde{m}_{\nu}(r\to\infty) that varies with respect to m~R\tilde{m}_{R}.

  • With this curve, the code solves the equation m~ν(r)=mν\tilde{m}_{\nu}(r\to\infty)=m_{\nu} with respect to m~R\tilde{m}_{R}. Depending on m¯ϕ\overline{m}_{\phi} and pF0p_{F0}, there can be two, one, or zero solutions. In the case of two solutions, one of them appears on the UR branch and the other appears on the NR branch.

Appendix H Annihilation cross sections

In this appendix, we present the detailed calculations of the annihilation cross sections. First, we need to compute the squared amplitude. For ν+ν¯\nu+\overline{\nu} annihilation, it is computed as follows:

||2\displaystyle|{\cal M}|^{2} =y4|v2¯[imν+imν]u1|2\displaystyle=y^{4}|\overline{v_{2}}\left[\frac{i}{\not{q}-m_{\nu}}+\frac{i}{\not{q}^{\prime}-m_{\nu}}\right]u_{1}|^{2}
=y4tr[(2mν)(1mν+1mν)(1+mν)(1mν+1mν)]\displaystyle=y^{4}{\rm tr}\left[\left(\not{p}_{2}-m_{\nu}\right)\left(\frac{1}{\not{q}-m_{\nu}}+\frac{1}{\not{q}^{\prime}-m_{\nu}}\right)\left(\not{p}_{1}+m_{\nu}\right)\left(\frac{1}{\not{q}-m_{\nu}}+\frac{1}{\not{q}^{\prime}-m_{\nu}}\right)\right]
=2y432mν8+16mν6(t+u)+mν4(t2+30tu+u2)mν2(t+u)(t2+14tu+u2)+tu(tu)2(mν2t)2(mν2u)2,\displaystyle=2y^{4}\frac{-32m_{\nu}^{8}+16m_{\nu}^{6}(t+u)+m_{\nu}^{4}\left(t^{2}+30tu+u^{2}\right)-m_{\nu}^{2}(t+u)\left(t^{2}+14tu+u^{2}\right)+tu(t-u)^{2}}{\left(m_{\nu}^{2}-t\right)^{2}\left(m_{\nu}^{2}-u\right)^{2}}, (157)

where p1p_{1} and p2p_{2} are the two initial momenta, qq and qq^{\prime} are the momenta of tt- and uu-channel neutrino propagators (t=qqt=q\cdot q and u=qqu=q^{\prime}\cdot q^{\prime}). They are related to the two final momenta p3p_{3} and p4p_{4} by q=p1p3q=p_{1}-p_{3} and q=p1p4q^{\prime}=p_{1}-p_{4}. In the second row, we have used v2v¯2=2mνv_{2}\overline{v}_{2}=\not{p}_{2}-m_{\nu}, u1u¯1=1+mνu_{1}\overline{u}_{1}=\not{p}_{1}+m_{\nu}.

In the non-relativistic limit and in the center-of-mass frame, we have

t=mν2+2cθΔΔ2+mν22Δ2,u=mν22cθΔΔ2+mν22Δ2,t=-m_{\nu}^{2}+2c_{\theta}\Delta\sqrt{\Delta^{2}+m_{\nu}^{2}}-2\Delta^{2},u=-m_{\nu}^{2}-2c_{\theta}\Delta\sqrt{\Delta^{2}+m_{\nu}^{2}}-2\Delta^{2}, (158)

where Δ=|𝐩1|=|𝐩2|\Delta=|\mathbf{p}_{1}|=|\mathbf{p}_{2}| and cθc_{\theta} is the cosine of the angle between 𝐩1\mathbf{p}_{1} and 𝐩3\mathbf{p}_{3}.

Plugging Eq. (158) to Eq. (157) and expanding it in terms of Δ\Delta, we obtain

||2=8y443cθ2mν2Δ2+𝒪(Δ4mν4).|{\cal M}|^{2}=8y^{4}\frac{4-3c_{\theta}^{2}}{m_{\nu}^{2}}\Delta^{2}+{\cal O}\left(\frac{\Delta^{4}}{m_{\nu}^{4}}\right). (159)

For ν+ν\nu+\nu annihilation, the calculation is similar:

||2\displaystyle|{\cal M}|^{2} =y4|u2¯[imν+imν]u1|2\displaystyle=y^{4}|\overline{u_{2}}\left[\frac{i}{\not{q}-m_{\nu}}+\frac{i}{\not{q}^{\prime}-m_{\nu}}\right]u_{1}|^{2}
=y4tr[(2+mν)(1mν+1mν)(1+mν)(1mν+1mν)]\displaystyle=y^{4}{\rm tr}\left[\left(\not{p}_{2}+m_{\nu}\right)\left(\frac{1}{\not{q}-m_{\nu}}+\frac{1}{\not{q}^{\prime}-m_{\nu}}\right)\left(\not{p}_{1}+m_{\nu}\right)\left(\frac{1}{\not{q}-m_{\nu}}+\frac{1}{\not{q}^{\prime}-m_{\nu}}\right)\right]
=2y48mν44mν2(t+u)+(tu)2(mν2t)(mν2u).\displaystyle=2y^{4}\frac{8m_{\nu}^{4}-4m_{\nu}^{2}(t+u)+(t-u)^{2}}{\left(m_{\nu}^{2}-t\right)\left(m_{\nu}^{2}-u\right)}. (160)

The main difference is that v2¯\overline{v_{2}} has been replaced by u2¯\overline{u_{2}}.

Plugging Eq. (158) to Eq. (160) and expanding it in terms of Δ\Delta, we obtain

||2=8y4(1+2cθ21mν2Δ2)+𝒪(Δ4mν4).|{\cal M}|^{2}=8y^{4}\left(1+\frac{2c_{\theta}^{2}-1}{m_{\nu}^{2}}\Delta^{2}\right)+{\cal O}\left(\frac{\Delta^{4}}{m_{\nu}^{4}}\right). (161)

In summary, the squared amplitude in the non-relativistic regime reads:

||2{8y4(43cos2θ)(v2)2for ν+ν¯ϕ+ϕ8y4for ν+νϕ+ϕ.|{\cal M}|^{2}\approx\begin{cases}8y^{4}\left(4-3\cos^{2}\theta\right)\left(\frac{v}{2}\right)^{2}&\text{for }\nu+\overline{\nu}\rightarrow\phi+\phi\\ 8y^{4}&\text{for }\nu+\nu\rightarrow\phi+\phi\end{cases}. (162)

With the above result of ||2|{\cal M}|^{2} , the total cross section of annihilation can be computed by

σ=164π[(p1p2)2mν4]||2𝑑q2.\sigma=\frac{1}{64\pi\left[\left(p_{1}\cdot p_{2}\right)^{2}-m_{\nu}^{4}\right]}\int|{\cal M}|^{2}dq^{2}. (163)

Here q2=(p3p1)2mν2(1+vcθq^{2}=(p_{3}-p_{1})^{2}\approx m_{\nu}^{2}(-1+vc_{\theta} varies in the range mν2(1v)q2mν2(1+v)m_{\nu}^{2}(-1-v)\lesssim q^{2}\lesssim m_{\nu}^{2}(-1+v). Integrating q2q^{2} in the kinetically allowed domain, we obtain the result in Eqs. (169) and (170).

Appendix I Energy loss due to ϕ\phi bremsstrahlung

Due to ϕ\phi bremsstrahlung and neutrino annihilation, the energy loss can cause further contraction of the virialized halo. Let us first consider the bremsstrahlung process in two-neutrino collision: ν+νν+ν+ϕ\nu+\nu\rightarrow\nu+\nu+\phi with ϕ\phi exchange in tt channel. The differential cross-section can be estimated as

dσdq2y28π2y416πv|q2|2,\frac{d\sigma}{dq^{2}}\sim\frac{y^{2}}{8\pi^{2}}\frac{y^{4}}{16\pi v|q^{2}|^{2}}\thinspace, (164)

where qq is the momentum carried by the tt-channel ϕ\phi mediator and vv is the relative velocity of two colliding neutrinos. The energy taken away by ϕ\phi equals roughly Eϕ|q2|/2mνE_{\phi}\simeq|q^{2}|/2m_{\nu}, so that according to (164) the energy loss is proportional to

σEϕy28π2qmin2qmax2y416πv(|q2|+mϕ2)2|q2|2mν𝑑q2y6256π3vmνlog(qmax2/mϕ2).\langle\sigma E_{\phi}\rangle\simeq\frac{y^{2}}{8\pi^{2}}\int^{q^{2}_{\max}}_{q^{2}_{\min}}\frac{y^{4}}{16\pi v(|q^{2}|+m_{\phi}^{2})^{2}}\frac{|q^{2}|}{2m_{\nu}}dq^{2}\simeq\frac{y^{6}}{256\pi^{3}vm_{\nu}}\log\left(q_{{\rm max}}^{2}/m_{\phi}^{2}\right). (165)

Here qmax2q_{{\rm max}}^{2} and qmin2q_{{\rm min}}^{2} are the maximal and minimal values of |q2||q^{2}|. The former is determined by kinematics, qmax2(mνv)2q_{{\rm max}}^{2}\sim(m_{\nu}v)^{2}, and for the latter we took qmin20q_{{\rm min}}^{2}\approx 0. Consequently, the rate of energy loss of a given neutrino can be estimated as

dElossdtσEϕnv.\frac{dE_{{\rm loss}}}{dt}\sim\langle\sigma E_{\phi}\rangle nv\thinspace. (166)

If q2rmean2q^{2}\ll r_{{\rm mean}}^{-2}, a given neutrino interacts coherently with neutrinos within the radius

rcoherent1/p,r_{{\rm coherent}}\sim 1/p\thinspace,

and the differential cross section can be enhanced by a factor of N~2\tilde{N}^{2}, where N~43πnrcoherent3\tilde{N}\sim\frac{4}{3}\pi nr_{{\rm coherent}}^{3} is the number of neutrinos within the coherence volume. On the other hand, taking N~\tilde{N} neutrinos as a target, the effective number density of scatterers is reduced to n/N~n/\tilde{N}. As a result, the rate in Eq. (166) still increases by a factor of N~\tilde{N}

N~=𝒪(10).\tilde{N}={\cal O}(10).

The energy loss rate per neutrino reads

dElossdty6N~n256π3mνlog(qmax2/mϕ2).\frac{dE_{{\rm loss}}}{dt}\sim\frac{y^{6}\tilde{N}n}{256\pi^{3}m_{\nu}}\log\left(q_{\rm max}^{2}/m_{\phi}^{2}\right). (167)

The time it takes for the system to cool down and lose half of its kinetic energy can be estimated by

τcoolingEK(vir)2N~(dEloss/dt),\tau_{{\rm cooling}}\sim\frac{E_{K}^{({\rm vir})}}{2\tilde{N}(dE_{{\rm loss}}/dt})\thinspace, (168)

where EKvirE_{K}^{\rm vir} is determined in Eq. (95). We take the virialized radius RvirRin/2R_{{\rm vir}}\sim R_{\rm in}/2 where RinR_{\rm in} is given Eq. (55). Then using n=N/(43πRvir3)n=N/(\frac{4}{3}\pi R_{{\rm vir}}^{3}) and v2=EK(vir)/(12mνN)v^{2}=E_{K}^{({\rm vir})}/(\frac{1}{2}m_{\nu}N), we obtain

τcooling1034sec1N~(107y)3(0.1eVmν)(pF00.1mν)5/2,\tau_{{\rm cooling}}\sim 10^{34}{\rm sec}\frac{1}{\tilde{N}}\left(\frac{10^{-7}}{y}\right)^{3}\left(\frac{0.1\ {\rm eV}}{m_{\nu}}\right)\left(\frac{p_{F0}}{0.1m_{\nu}}\right)^{-5/2},

which is much larger than the age of the Universe τuniverse=4.35×1017\tau_{\text{universe}}=4.35\times 10^{17} sec.

Appendix J Annihilation

Neutrino annihilation rate depends on whether neutrinos are Dirac or Majorana particles. For Dirac neutrinos, the annihilation channel is ν+ν¯ϕ+ϕ\nu+\overline{\nu}\rightarrow\phi+\phi (see the second diagram in Fig. 8). For the Majorana neutrinos, there is an additional process ν+νϕ+ϕ\nu+\nu\rightarrow\phi+\phi with Majorana mass insertion (see the first diagram in Fig. 8 ).

Refer to caption
Figure 8: Majorana and Dirac neutrino annihilation.

In the non-relativistic case the cross sections read (see Appendix H for details of the calculation):

σνν¯σ(ν+ν¯ϕ+ϕ)\displaystyle\sigma^{\nu\bar{\nu}}\equiv\sigma(\nu+\overline{\nu}\rightarrow\phi+\phi) 3vy416πmν2,\displaystyle\approx\frac{3vy^{4}}{16\pi m_{\nu}^{2}}\thinspace, (169)
σννσ(ν+νϕ+ϕ)\displaystyle\sigma^{\nu\nu}\equiv\sigma(\nu+\nu\rightarrow\phi+\phi) y44πmν2v,\displaystyle\approx\frac{y^{4}}{4\pi m_{\nu}^{2}v}\thinspace, (170)

where vv is the relative velocity of colliding neutrinos:

v=v12+v222v1v2cosθ12,v=\sqrt{v_{1}^{2}+v_{2}^{2}-2v_{1}v_{2}\cos\theta_{12}}\thinspace,

and vi=|𝐩i|/mνv_{i}=|\mathbf{p}_{i}|/m_{\nu} (i=1,2i=1,2) are the velocities of the colliding neutrinos and θ12\theta_{12} the angle between 𝐩1\mathbf{p}_{1} and 𝐩2\mathbf{p}_{2}.

The key difference is dependence on vv: σνν¯v\sigma^{\nu\bar{\nu}}\propto v, while σννv1\sigma^{\nu\nu}\propto v^{-1}. For the Majorana case (both ν+ν¯\nu+\overline{\nu} and ν+ν\nu+\nu channels are present) in the limit v1v\ll 1, the left diagram dominates. Hence for Dirac or Majorana neutrinos, one only needs to consider ν+ν¯\nu+\overline{\nu} or ν+ν\nu+\nu annihilation, respectively.

The change of number density due to annihilation can be computed as

dndt=f1(𝐩1)f2(𝐩2)σ(v)vd3𝐩1(2π)3d3𝐩2(2π)3,\frac{dn}{dt}=-\int f_{1}(\mathbf{p}_{1})f_{2}(\mathbf{p}_{2})\sigma(v)v\frac{d^{3}\mathbf{p}_{1}}{(2\pi)^{3}}\frac{d^{3}\mathbf{p}_{2}}{(2\pi)^{3}}\thinspace, (171)

where f1f_{1} and f2f_{2} are the distribution functions of the initial neutrinos. Introducing integration over velocities the rate (171) becomes

dndt=f1f2σ(v)vmν34πv12dv1(2π)3mν32πv22dv2dcosθ12(2π)3,\frac{dn}{dt}=-\int f_{1}f_{2}\sigma(v)v\frac{m_{\nu}^{3}4\pi v_{1}^{2}dv_{1}}{(2\pi)^{3}}\frac{m_{\nu}^{3}2\pi v_{2}^{2}dv_{2}d\cos\theta_{12}}{(2\pi)^{3}}\thinspace,

The result reads

dnνν¯dt\displaystyle\frac{dn^{\nu\bar{\nu}}}{dt} pF8y4160π5mν4,dnννdtpF6y4144π5mν2.\displaystyle\approx-\frac{p_{F}^{8}y^{4}}{160\pi^{5}m_{\nu}^{4}},\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \frac{dn^{\nu\nu}}{dt}\approx-\frac{p_{F}^{6}y^{4}}{144\pi^{5}m_{\nu}^{2}}.\ (172)

Using Eq. (31), we express the number density nn in terms of pFp_{F}, thus obtaining the differential equation for pFp_{F}:

dpFνν¯dt\displaystyle\frac{dp_{F}^{\nu\bar{\nu}}}{dt} =pF6y480π3mν4,\displaystyle=-\frac{p_{F}^{6}y^{4}}{80\pi^{3}m_{\nu}^{4}}\,, (173)
dpFννdt\displaystyle\frac{dp_{F}^{\nu\nu}}{dt} =pF4y472π3mν2,\displaystyle=-\frac{p_{F}^{4}y^{4}}{72\pi^{3}m_{\nu}^{2}}\,, (174)

The solution is

pFνν¯(t)\displaystyle p_{F}^{\nu\bar{\nu}}(t) =pF0[1+pF05y416π3mν4t]1/5,\displaystyle=p_{F}^{0}\left[1+\frac{p_{F0}^{5}y^{4}}{16\pi^{3}m_{\nu}^{4}}t\right]^{-1/5}\,, (175)
pFνν(t)\displaystyle p_{F}^{\nu\nu}(t) =pF0[1+pF03y424π3mν2t]1/3,\displaystyle=p_{F}^{0}\left[1+\frac{p_{F0}^{3}y^{4}}{24\pi^{3}m_{\nu}^{2}}t\right]^{-1/3}\,, (176)

where pF0p_{F}^{0} is the Fermi momentum in the beginning of the cooling phase. The annihilation lifetime τanni\tau_{{\rm anni}} can be defined as the time during which the Fermi momentum reduces by a factor of 2: pF(τanni)=pF0/2p_{F}(\tau_{{\rm anni}})=p_{F}^{0}/2, which corresponds to the density being reduced by a factor of 8. Then from (175) and (176) we obtain

τνν¯\displaystyle\tau^{\nu\bar{\nu}} =496π3mν4y4pF05=1.0×1023seconds×(0.1eVmν)(0.1mνpF0)5(107y)4,\displaystyle=\frac{496\pi^{3}m_{\nu}^{4}}{y^{4}p_{F0}^{5}}=1.0\times 10^{23}\text{seconds}\times\left(\frac{0.1\ {\rm eV}}{m_{\nu}}\right)\left(\frac{0.1m_{\nu}}{p_{F0}}\right)^{5}\left(\frac{10^{-7}}{y}\right)^{4}, (177)
τνν\displaystyle\tau^{\nu\nu} =168π3mν2y4pF03=3.4×1020seconds×(0.1eVmν)(0.1mνpF0)3(107y)4.\displaystyle=\frac{168\pi^{3}m_{\nu}^{2}}{y^{4}p_{F0}^{3}}=3.4\times 10^{20}\text{seconds}\times\left(\frac{0.1\ {\rm eV}}{m_{\nu}}\right)\left(\frac{0.1m_{\nu}}{p_{F0}}\right)^{3}\left(\frac{10^{-7}}{y}\right)^{4}. (178)

Therefore for typical values of parameters presented in Eqs. (177) and (178), one finds that that τanniτuniverse=4.35×1017seconds\tau_{{\rm anni}}\gg\tau_{\text{universe}}=4.35\times 10^{17}{\rm seconds}.

Acknowledgements.
The authors would like to thank Julian Heeck and Arvind Rajaraman for useful discussions on this topic in 2019, Jerome Vandecasteele and Peter Tinyakov for discussions on similar fermionic systems in neutron stars, and Laura Lopez Honorez for suggesting references. X.J.X is supported in part by the National Natural Science Foundation of China under grant No. 12141501.

References