1. Introduction
Neutrino oscillation is a quantum mechanical phenomenon in which neutrinos change their flavor (electron, muon, or tau) as they propagate through space. This behavior demonstrates that neutrinos have mass, a fact that challenges the Standard Model of particle physics, which initially assumed massless neutrinos. The discovery of neutrino oscillations was first conclusively demonstrated through the observation of atmospheric neutrinos by the Super–Kamiokande experiment. In 1998, the Super–Kamiokande collaboration reported a deficit in the expected number of muon neutrinos arriving from the atmosphere, depending on the distance traveled, providing strong evidence for oscillation and therefore non-zero neutrino mass [
1]. This groundbreaking result has profound implications: it not only necessitates an extension of the Standard Model but also opens new avenues in cosmology and particle physics, including the role of neutrinos in the evolution of the universe and the search for CP violation in the lepton sector, among other areas.
The confirmation that neutrinos have mass and oscillate between flavors has significantly deepened our understanding of their role in astrophysical environments, especially in extreme conditions such as those found in neutron stars and magnetars. These ultra-dense stellar remnants, particularly magnetars, whose magnetic fields can reach surface intensities of 10
14–10
15 G and increase dramatically within their dense interiors [
2], provide natural laboratories for studying neutrino behavior under intense magnetic forces. In such environments, neutrino emission becomes a dominant cooling mechanism and influences the star’s thermal and magnetic evolution [
3]. Within the broader context of core-collapse supernovae, the review by Burrows and Vartanyan [
4] provides a comprehensive overview of the problem, addressing not only neutrino physics but also the evolution of the supernova remnant. Moreover, the strong magnetic fields in magnetars can affect neutrino propagation and oscillation patterns, potentially leading to flavor conversions enhanced by matter and magnetic effects [
5,
6,
7,
8]. Understanding neutrino interactions in these settings is crucial not only for modeling supernovae and neutron star formation but also for probing the physics of dense matter, magnetic field generation, and the potential existence of sterile neutrinos or non-standard interactions beyond the Standard Model [
9]. Thus, the study of neutrino oscillations not only reshapes fundamental particle physics but also enriches our comprehension of the most violent and enigmatic objects in the cosmos.
One particularly intriguing astrophysical phenomenon potentially linked to neutrino behavior in highly magnetized neutron stars is the so-called pulsar kick: the observation that many pulsars are born with high space velocities, often exceeding
, far greater than those of their progenitor stars [
10]. While several mechanisms have been proposed, one compelling explanation involves anisotropic neutrino emission during the proto-neutron star cooling phase, especially in the presence of ultra-strong magnetic fields. These fields can polarize the medium, thus modifying the cross-sections for neutrino interactions, resulting in asymmetric momentum transfer that imparts a net recoil to the newly formed neutron star [
11]. Magnetars, with their extreme field strengths, provide a natural context in which such asymmetries could be amplified. Furthermore, certain models suggest that parity-violating weak interactions in a magnetized medium could generate directional neutrino fluxes aligned with the magnetic axis [
12]. These processes offer a viable channel for generating the observed natal kicks and connect large-scale neutron star dynamics with microphysical processes occurring under extreme conditions, thus linking compact object astrophysics with fundamental particle interactions.
In this work, we focus on the inelastic scattering of neutrinos with neutrons, a mechanism responsible for the asymmetric emission of neutrinos. More specifically, we evaluate the neutrino mean free path (defined as the inverse of the total neutrino cross section per unit volume) in hot, dense matter under the influence of a strong magnetic field. The neutrino mass is very small, eV, and this non-zero mass allows for the presence of a right-handed component in the neutrino state, although its contribution is expected to be negligible. In the limiting case of massless neutrinos, only fully polarized left-handed neutrinos exist. These represent the dominant contribution and are the main focus of this work. Nevertheless, to gain some insight into the physics of right-handed neutrinos, we also discuss this point.
It is worth mentioning that we have previously studied the mean free path of left-handed neutrinos in Ref. [
13], where a specific set of approximations was employed. In the present work, we revisit these approximations, which leads to the emergence of quantum interference terms that were not considered before.
The neutrino–neutron inelastic scattering cross section can be evaluated either in free space or within a dense medium. However, the mean free path is physically meaningful only in a dense medium, which must be characterized by an equation of state (EoS). In this work, we explore three different equations of state. The first one assumes no strong interaction among neutrons, whereas the remaining two are based on the Hartree–Fock approximation employing the Skyrme interaction, commonly known as the Skyrme model.
The neutrino mean free path in the absence of a magnetic field has been studied by many authors using various approximation schemes and models of the trapping environment (see, e.g., Refs. [
14,
15,
16,
17,
18,
19,
20,
21,
22,
23,
24,
25,
26,
27], and references therein). The behavior of neutrinos in dense matter under strong magnetic fields has also been explored in the literature [
13,
28,
29,
30,
31,
32,
33,
34,
35,
36,
37,
38,
39,
40]. However, the asymmetry in neutrino emission caused by the breaking of isotropy due to the magnetic field has not been extensively discussed.
This paper is organized as follows: In
Section 2, we briefly discuss neutrino polarization. In
Section 3, we present elements of the EoS and derive an analytical expression for the neutrino cross section per unit volume for left-handed neutrinos. In
Section 4, we discuss our results, and finally, in
Section 5, we present our conclusions.
2. Neutrino Polarization: A Brief Discussion
The existence of a nonzero neutrino mass has profound implications for the structure of neutrino states. In the massless limit, neutrinos are purely left-handed and antineutrinos are right-handed, as dictated by the Standard Model. However, once a mass term is introduced, each neutrino state acquires both left- and right-helicity components due to the Lorentz structure of the Dirac spinor. The right-helicity component of a neutrino (or left-helicity component of an antineutrino) becomes nonzero, but its amplitude is suppressed by a factor proportional to , where is the neutrino mass and its energy. Given the extremely small mass of the neutrino compared to typical energies involved, the right-helicity component is extremely suppressed and often considered unobservable. Nonetheless, its existence is a direct consequence of neutrino mass and can, in principle, be probed in precision experiments involving polarized sources or detectors.
In principle, the presence of both helicity components in a massive neutrino state opens the possibility of helicity oscillations, especially when neutrinos propagate through external fields or media. These transitions between left- and right-helicity states are suppressed in vacuum due to the smallness of the mass, but can be enhanced in certain conditions, such as in the presence of strong magnetic fields or dense matter through mechanisms involving spin-flip processes or magnetic moment interactions [
41,
42,
43,
44]. Despite their tiny amplitude, right-helicity neutrino components could have significant implications in astrophysical environments, where long propagation distances and extreme conditions might lead to cumulative effects. Moreover, if right-handed neutrinos are sterile, i.e., non-interacting under the Standard Model forces, the conversion of active left-handed neutrinos into sterile right-handed states could impact neutrino fluxes observed in experiments and cosmological observables. Thus, even a seemingly negligible component may play a crucial role in scenarios beyond the Standard Model.
Unfortunately, the interaction between right-handed neutrinos and neutrons is not known. Within the framework of the Standard Model, such an interaction is strictly absent, implying a vanishing cross section for right-handed neutrinos. However, as we shall discuss below, the presence of a mass term leads to energy eigenstates that are admixtures of left- and right-handed components. This mixing induces a small correction to the left-handed component, even in states initially produced as purely left-handed. As a result, the modified wave function can exhibit a suppressed but nonzero sensitivity to phenomena that would otherwise involve only right-handed neutrinos. This subtle effect, though minute, forms the basis for exploring possible deviations from Standard Model predictions in neutrino scattering and propagation.
In a more specific way, if the mass of the neutrino is nonzero, then the energy eigenstates are not eigenstates of chirality. The energy eigenstates can be written as a linear combination of left- and right-chiral states through a unitary transformation:
where
is the positive (negative) energy eigenstate, and
denotes the left- (right-) chiral component of the neutrino.
In the limiting case of massless neutrinos,
, and we recover the following:
as expected, since helicity and chirality coincide for massless fermions.
The mixing angle
is given by the following:
Let us focus on Equation (
1). The right-chiral component corresponds to a sterile neutrino, whose interactions with a neutron (or other Standard Model particles) are highly suppressed or unknown. Therefore, when calculating the neutrino–neutron cross section, only the left-chiral component contributes. However, due to the nonzero neutrino mass, the left-chiral component is multiplied by a factor
, reflecting the admixture of the sterile (right-chiral) component. The importance of this correction depends directly on the energy of the neutrino: the smaller the energy, the larger the deviation from a pure left-chiral state.
4. Results and Discussion
In this section, we present our results for the neutrino and antineutrino mean free paths in hot, dense neutron matter under a strong magnetic field. We analyze a density range of
fm
−3, several temperatures up to
MeV, and various magnetic field intensities ranging from
to
G. To describe the medium, we employ different models for the Equation of State (EoS) in order to assess the sensitivity of the mean free path to the properties of the medium. Most of the results are obtained using the LNS Skyrme interaction developed by Cao et al. [
52], which we adopt because its particular density dependence of the effective mass makes it especially suitable for our framework. Unless otherwise specified, the equation of state just detailed is to be taken as the default model.
In all cases, neutrinos and antineutrinos are assumed to be massless and not trapped in the medium, so their chemical potential vanishes (
). They are described by the standard Fermi–Dirac distribution,
with energy
. Since
, neutrinos and antineutrinos share the same distribution and energy for a given momentum. A representative value for their energy in a thermal medium is the average,
which corresponds to the thermal average energy of a relativistic fermion in equilibrium. However, for the sake of practicality and in line with common choices in the literature, we adopt the approximate value
in the following calculations. An immediate consequence of this approximation, considering that we explore temperatures starting from
, is that the effect of the neutrino mass is negligible: the ratio
. It is worth noting that it is also possible to work at
, but in that case, the neutrino energy must be modeled according to the emission mechanism. This conclusion is independent of the EoS, as it concerns the neutrino wave function, which does not enter into the EoS description.
Before discussing our results, it is worth noting that, for trapped neutrinos and antineutrinos, their distribution functions differ as follows:
This implies that
, which would in turn lead to different mean free paths for neutrinos and antineutrinos. In addition, due to the definition of helicity, left-handed neutrinos and right-handed antineutrinos share the same analytical expression for the spinor (see Equations (
A7) and (
A9), respectively). Therefore, within our model, neutrinos and antineutrinos have identical mean free paths.
To begin the discussion of our results, we compare them with those previously obtained in Ref. [
13], where certain approximations were made. Specifically, the neutron in the initial state was assumed to have either spin up or spin down. Furthermore, it was also assumed that the neutron’s final state always retained the same spin orientation as the initial state. In our current model, the initial neutron spin state, given in Equation (
A4), is a mixed state in which each spin component is weighted by a factor. These factors are chosen based on the results of the equation of state, so as to match the mean-field spin projection with the system’s polarization. The use of such a wave function introduces quantum interference terms that allow for the initial and final neutron spin states to differ.
In
Figure 1, we show the previous results (dashed lines) and compare them with those obtained using our improved scheme (solid lines). We recall that the
-axis is defined along the direction of the constant magnetic field, thereby establishing a preferred spatial direction. In the absence of a magnetic field, the only relevant angle is the relative angle between the incoming and outgoing neutrino. In the presence of the field, however, we must consider the angle
between the initial neutrino direction and the magnetic field, the corresponding
for the outgoing neutrino, and the azimuthal angle
of the outgoing neutrino. The weak interaction dynamics lead to different values for the neutrino mean free path depending on
, as shown in
Figure 1. This asymmetry in the mean free path suggests that, due to the scattering process, a higher number of neutrinos are expected to travel in the direction of the magnetic field, where the mean free path is longer.
Continuing with the discussion of
Figure 1, we observe good consistency between both approximations. At low densities, the results for
are very similar. For
and
, our present model shows a reduction in the asymmetry. The situation changes at higher densities: for all angles, the mean free path is longer in the present model. In any case, the differences between both models are noticeable. Two points should be emphasized: first, in Ref. [
13], we noted that the results for
were nearly identical for
and
. This agreement remains valid in our improved model at low densities, but it no longer holds at high densities, where the mean free path increases compared to the previous model. Second, in this figure we have chosen a temperature of
MeV. This temperature smooths the results. We will later discuss the case of
MeV, which requires a more detailed analysis.
To enable a more quantitative analysis of these results, we define the function
where
denotes the neutrino mean free path obtained using the previous approximations from Ref. [
13], and
corresponds to the result from our improved model. Note that the mean free path depends functionally on
,
T,
B, and
, i.e.,
, although this dependence is omitted for brevity.
Table 1 presents the values of the function
for several densities and neutrino incident angles. This table supports the qualitative discussion given in
Figure 1. Beyond the consistency between the two approximations, the differences introduced by the improved model are significant enough to justify its implementation. In particular, the value
for
and
stands out. This result has important implications within our framework, although a detailed discussion is deferred to a later stage, specifically when analyzing the case of temperature
MeV. It is well understood that all models are approximations and inherently include some form of quantum interference terms. While our proposed model provides a more accurate treatment of certain quantum interference effects, direct comparisons between models remain challenging. Throughout this work, we somewhat arbitrarily use the term “interference term” as a shorthand to refer to our specific treatment of these terms.
In what follows, we study the behavior of the neutrino mean free path as a function of the magnetic field strength, temperature, and the equation of state. We begin in
Figure 2, showing results for different magnetic field strengths ranging from
G up to
G. For
and at low densities, the results are nearly identical for all values of
B (including
). At high densities, we observe the previously mentioned increase in the mean free path for
G, which decreases with lower values of
B and almost vanishes at
G. The asymmetry in the neutrino mean free path is significant for
G, barely noticeable for
G, and very small for the remaining values.
Next, in
Figure 3, we analyze the effect of temperature on the neutrino mean free path. Recall that we evaluate the mean free path as the inverse of the total cross section per unit volume, as given by Equation (
23). The available phase space is determined by the distribution functions
(see Equation (
8)). As the temperature increases, these distribution functions allow access to a larger phase space, leading to an increase in the cross section and, consequently, a decrease in the mean free path, as shown in
Figure 3. Regarding the asymmetry in the mean free path, it also decreases with increasing temperature (note the logarithmic scale in the figure). This is due to the reduction in spin polarization of the system at higher temperatures, which in turn leads to a reduced asymmetry in the mean free path. Thermal disorder tends to suppress spin alignment. Beyond these considerations, note the crossing of the mean free paths for
and
at
MeV. This curious result is discussed in detail at the end of this section.
We now turn to the analysis of the dependence of the neutrino mean free path on different models for the equation of state (EoS). This study is performed using three EoS models: one that includes no strong interactions between neutrons, and two based on different parameterizations of the Skyrme-type effective interaction. The first is the LNS Skyrme model, and the second is the SLy4 model developed by Douchin et al. [
53]. The Skyrme framework incorporates strong interactions at the mean-field level. In the absence of strong interactions, the EoS is determined solely by phase-space considerations arising from the Pauli exclusion principle. These three EoS models have been selected for the following reasons. The parameters of the LNS Skyrme model were fitted to reproduce the nuclear matter EoS obtained in the non-relativistic Brueckner–Hartree–Fock (BHF) approach using the Argonne V18 [
54] nucleon–nucleon potential, supplemented with the Urbana IX [
55] three-nucleon force. This model yields neutron effective masses that do not deviate significantly from the bare mass; however, it exhibits a so-called ferromagnetic instability at high densities. In contrast, the SLy4 model does not present this issue, but it predicts small effective masses, which, as we will show, leads to an unexpected increase in the neutrino mean free path. Finally, the inclusion of an EoS without strong interactions serves to isolate and highlight the role of strong interactions in determining the neutrino mean free path. A similar analysis for charge-exchange reactions was presented in [
56].
For convenience, we discuss this topic using two figures,
Figure 4 and
Figure 5, corresponding to
MeV and
MeV, respectively. In both figures, panel (a) corresponds to the free (non-interacting) EoS, panel (b) to the LNS Skyrme model, and panel (c) to the Douchin parameterization of the strong interaction. From both figures, our first conclusion is that the neutrino mean free path shows a strong dependence on the EoS, as the results differ significantly across the various models. We emphasize this point, as our calculation is fully self-consistent: we compute the EoS, and from it, determine the spin asymmetry, the chemical potential, and the single-particle energies, which are then used in the calculation of the neutrino mean free path. To the best of our knowledge, there is no universally preferred set of EoS parameters. As a result, the predictions for the neutrino mean free path remain model-dependent, pending observational data that could provide constraints to select the most appropriate EoS.
We now examine the behavior of the neutrino mean free path in each panel individually. For simplicity, we focus on
Figure 4 (the discussion for
Figure 5 is very similar). The results for the free EoS shown in panel (a) illustrate the basic concepts of the problem: as the density increases, the probability of scattering events increases, and thus the mean free path decreases. Regarding the system’s spin asymmetry, increasing the baryon density enhances the effect of the Pauli exclusion principle, favoring a more symmetric configuration with equal numbers of spin-up and spin-down particles. A direct consequence of this reduced spin asymmetry is a decrease in the asymmetry of the neutrino mean free path, as seen in panel (a).
Panels (b) and (c) include the effects of the strong interaction, leading to a more complex behavior. At low densities, the (short-range) strong interaction plays a minor role, and the behavior resembles that of panel (a), albeit with a slightly reduced asymmetry in the mean free path. At higher densities, the strong interaction becomes more significant, and an increase in the mean free path with density is observed. This increase arises from two different effects. In particular, in panel (b), there is a non-physical magnetization of the system induced by the Skyrme interaction. Additionally, at high densities, a pronounced reduction in the neutron effective mass contributes to the increased mean free path; this effect is especially important in panel (c).
In both
Figure 4 and
Figure 5, we observe an increase in the neutrino mean free path for
relative to the two remaining angles. This effect is subtle at
MeV but becomes very pronounced at
MeV, where, in all panels, a crossing occurs with the result for
. The origin of this behavior will be discussed in detail in the remainder of this section. From these figures, our first conclusion is that this feature is unlikely to be directly related to the EoS, as it appears consistently across all models considered.
The origin of the crossing of the mean free paths discussed above is subtle and involves several elements. It arises from quantum interference terms, which, to the best of our knowledge, are analyzed for the first time in the present work. These terms result from the dynamics of the weak interaction. At this stage, and for the purpose of gaining insight into the underlying mechanism, it is more convenient to consider the total cross section per unit volume rather than the mean free path. For simplicity, we focus on the free interaction EoS at
MeV. The total cross section is given in Equation (
23) and can be naturally decomposed into four terms according to the spin projections of the initial and final neutron states:
,
,
, and
. For instance, the
term corresponds to an initial neutron with spin up and a final neutron with spin down. This decomposition is shown in
Figure 6. All curves in panels (a) and (b) exhibit the expected behavior, namely a monotonic increase with density. What ultimately determines the behavior of the total cross section is the relative contribution of each term.
Panel (a) shows the sum of the direct contributions () and the interference terms () for two relevant neutrino angles. All contributions increase monotonically with density. The direct terms converge at high densities, with taking consistently lower values for . The interference terms, on the other hand, behave differently: is larger for than for , and the difference between the two remains approximately constant at medium and high densities. This panel reveals that the total contribution () exhibits a crossing between the results for the two angles. This identifies the origin of the nontrivial behavior, though it does not yet constitute a full explanation. It is now clear that the interference terms are responsible for the observed feature. As already mentioned, to the best of our knowledge, these contributions are accounted for explicitly here for the first time.
We now turn to the trace function
in Equation (
20). From this expression, we define two functions associated with the interference contributions:
Inspection of Equation (
20) shows that
and
are associated with the
and
contributions, respectively.
Panel (b) of
Figure 6 shows the individual interference terms. It is useful to keep in mind the expressions for
and
, and their values for
and
, listed in
Table 2. From this panel, one finds that
which can be directly understood from
Table 2.
It remains to understand why
. This result is somewhat counterintuitive, given that the system’s polarization favors neutrons with spin down. In Equation (
20), we observe that the
contribution is weighted by the factor
, while the corresponding factor for
is
, with
. This inequality in the cross section can be understood in terms of the structure function
, defined in Equation (
24).
Figure 7 shows the structure function as a function of
for a representative value of
, selected from the numerical analysis.
Under the conditions of
Figure 7 (
MeV), the cross section receives contributions from values of
within the approximate range
MeV. These values are schematic in nature and are presented solely for illustrative purposes, to highlight the features of our results. Notably, over the entire range of
,
dominates over
. However, due to the fixed neutrino energy
, the kinematics restrict the range of accessible
values, as discussed above. This restriction accounts for the observed hierarchy between
and
and also explains why the effect becomes less pronounced at higher temperatures.
In summary, the peculiar behavior originates in the weak interaction dynamics, encoded in the functions
and
. As a final point, in
Figure 8, we present the neutrino mean free path as a function of the incoming neutrino angle
for two different baryon densities. At low density, the result is as expected: neutrinos are preferentially emitted parallel to the magnetic field. However, at high density, the maximum occurs near
. This outcome significantly alters the prevailing paradigm regarding asymmetric neutrino emission due to neutrino–neutron scattering. It indicates that at high densities and low temperatures, neutrinos are predominantly emitted perpendicular to the magnetic field. As discussed, this result is independent of the equation of state.
Before concluding this section, it is worth noting that, due to the scale of the reaction under consideration, the density, temperature, and magnetic field are treated as locally constant. In a realistic neutron star model, these quantities vary with position inside the star. To implement such a model, it is necessary to know the cross section (or, equivalently, the mean free path) for the reaction of interest, among other inputs; this is the focus of the present contribution. In a recent work [
56], we addressed charge-exchange reactions under the same level of approximation. Unfortunately, our non-relativistic model is not suitable for describing the dense stellar core; a relativistic treatment is necessary. Together with the inclusion of nuclear correlations beyond the mean-field level, we consider the development of a relativistic model a possible direction for future work, although it lies beyond the scope of the present study. Under the conditions considered here, a fully relativistic model should reproduce our results in the appropriate limit. This highlights the role of the non-relativistic result as a guiding reference for the more comprehensive relativistic treatment.
Regarding the pulsar kick problem, our present results disfavor asymmetric neutrino emission as its explanation. This stands in contrast to the findings for charge-exchange reactions discussed in Ref. [
56]. Ultimately, incorporating all relevant reactions into a realistic neutron star model would be necessary to resolve this issue conclusively. For a recent review of the pulsar kick problem, we refer the reader to the work of Lambiase and Poddar [
57], where several possible explanations are briefly discussed.