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

Analytical Expressions for Neutrino Oscillation

Adriano Cherchiglia Instituto de Física Gleb Wataghin, Universidade Estadual de Campinas
Rua Sérgio Buarque de Holanda, 777, Campinas, SP, Brasil
   Macello Jales Instituto de Física Gleb Wataghin, Universidade Estadual de Campinas
Rua Sérgio Buarque de Holanda, 777, Campinas, SP, Brasil
   Guilherme Nogueira Instituto de Física Gleb Wataghin, Universidade Estadual de Campinas
Rua Sérgio Buarque de Holanda, 777, Campinas, SP, Brasil
   Maressa P. Sampaio Instituto de Física Gleb Wataghin, Universidade Estadual de Campinas
Rua Sérgio Buarque de Holanda, 777, Campinas, SP, Brasil
   Pedro C. de Holanda Instituto de Física Gleb Wataghin, Universidade Estadual de Campinas
Rua Sérgio Buarque de Holanda, 777, Campinas, SP, Brasil
Abstract

Research in neutrino physics has been very active, both in experimental advances, with a new generation of detectors in operation and planning, and in theoretical discussions regarding the fundamental nature of the neutrino. This scientific dynamism has attracted many new students to the field. One of the first topics studied in neutrino physics by newcomers is the formalism of neutrino flavor oscillations and its associated phenomenology. We present this work as a compilation of this basic knowledge, through a step-by-step approach that facilitates an efficient understanding of this vast theoretical and experimental landscape.

I Introduction

The Standard Model of Particle Physics (SM) was firmly established in the 70’s, encompassing three of the four fundamental interactions observed in Nature. In this framework, apart from the gauge bosons related to the electromagnetic and strong interactions, there was only a particle considered massless: the neutrino. Contrary to the gauge bosons, whose massless nature is directly connected to the conserved symmetries of the SM, there was no a priori theoretical reason for massless neutrino. The main reason was phenomenological.

This picture remained untouched until the end of the 90’s and begin of 2000’s, when there was solid confirmation of the phenomenon of neutrino oscillation, both by Super-Kamiokande [1] and the Sudbury Neutrino Observatory (SNO)  [2]. This phenomenon can only be explained if neutrinos are massive particles, contradicting the hypothesis included in the SM. This fact can be regarded as one of the few solid experimental observations that require the SM to be extended.

A particular interesting question is how to accommodate a massive neutrino within the Higgs mechanism. This is only possible if neutrinos with right-handed chirality are introduced. However, since their electroweak hypercharge is null, they do not couple with the neither the photon or the Z boson. They also do not couple with the W boson, since charged currents are left-handed. Therefore, in the framework of the SM particles, they could only couple with the Higgs. Since neutrinos masses are tiny, this coupling would be also very small, extremely challenging to be detected by present and future experiments. Nevertheless, other mechanisms for generating neutrinos masses have been put forward, which require the inclusion of other particles beyond the SM (see, for instance, [3] for a review).

Apart from the mechanism to generate their masses, neutrinos hide other mysteries. In particular, they are the only particles in the SM that could be their own antiparticles (Majorana), since they do not carry electromagnetic (or colour) charge. On-going experiments on neutrinoless double beta decay aim to unveil this characteristic [4, 5]. Another intriguing question is whether CP violation can also occur in the leptonic sector. Current and near-future oscillation neutrino experiments aim to provide a definite answer for this question. At present, the CP conservation seems to be disfavored, but more date is needed.

Given this general picture, the study of neutrinos properties presents itself as a promising path towards an extended SM. Therefore, precise knowledge of the neutrino oscillation phenomenon is desirable, motivating next generation experiments such as DUNE, Hyper-Kamiokande and JUNO. In a nutshell, the differential event rate for neutrinos of flavor β\beta with energy EνE_{\nu} to be detected at a distance LL from the source SS, where they were produced with flavor α\alpha, is given by

RαβS=NTΦα(Eν)σβ(Eν)PαβR_{\alpha\beta}^{S}=N_{T}\Phi_{\alpha}(E_{\nu})\sigma_{\beta}(E_{\nu})P_{\alpha\beta} (1)

where NTN_{T} is the number of target particles, and Φα(Eν)\Phi_{\alpha}(E_{\nu}), σβ(Eν)\sigma_{\beta}(E_{\nu}) are the incident flux and detection cross-section, respectively. Particularly important are the neutrino conversion probabilities PαβP_{\alpha\beta}, which are the main topic of our work. We will review their definition in sec.II, providing analytic expressions both in vacuum and in the modifications of the neutrino parameters in matter (sec.III). A series of approximations will be discussed, as well as their limits of applicability. We aim to provide the reader with easy-to-use expressions, helping into developing their physical intuition regarding the phenomenon of neutrino oscillation.

II Vacuum Oscillation Probabilities

The usual expression for the neutrino flavor probabilities, evaluated within the framework of quantum mechanics, can be written as [6]

Pαβ=δαβ\displaystyle P_{\alpha\beta}=\delta_{\alpha\beta} \displaystyle- 4k>j[UαkUβkUαjUβj]sin2(Δmkj2L4E)\displaystyle 4\sum_{k>j}\Re\left[U^{*}_{\alpha k}U_{\beta k}U_{\alpha j}U^{*}_{\beta j}\right]\sin^{2}\left(\frac{\Delta m^{2}_{kj}L}{4E}\right) (2)
+\displaystyle+ 2k>j[UαkUβkUαjUβj]sin(Δmkj2L2E).\displaystyle 2\sum_{k>j}\Im\left[U^{*}_{\alpha k}U_{\beta k}U_{\alpha j}U^{*}_{\beta j}\right]\sin\left(\frac{\Delta m^{2}_{kj}L}{2E}\right).

The dependence on neutrino masses in Eq. (2) is encompassed by the difference in the squares of the individual mass eigenstates, Δmkj2=mk2mj2\Delta m_{kj}^{2}=m_{k}^{2}-m_{j}^{2}. Meanwhile, the ratio of the distance between the neutrino source and the detector, LL, by the neutrino energy EE is known as the baseline. Additionally, UU is the PMNS matrix, which in its common parametrization [7] takes the following form

U\displaystyle U =\displaystyle= R(θ23)R(θ13,δCP)R(θ12)=(c12c13s12c13s13eiδCPs12c23c12s23s13eiδCPc12c23s12s23s13eiδCPs23c13s12s23c12c23s13eiδCPc12s23s12c23s13eiδCPc23c13),\displaystyle R(\theta_{23})R(\theta_{13},\delta_{CP})R(\theta_{12})=\left(\begin{array}[]{ccc}c_{12}c_{13}&s_{12}c_{13}&s_{13}e^{-i\delta_{CP}}\\ -s_{12}c_{23}-c_{12}s_{23}s_{13}e^{i\delta_{CP}}&c_{12}c_{23}-s_{12}s_{23}s_{13}e^{i\delta_{CP}}&s_{23}c_{13}\\ s_{12}s_{23}-c_{12}c_{23}s_{13}e^{i\delta_{CP}}&-c_{12}s_{23}-s_{12}c_{23}s_{13}e^{i\delta_{CP}}&c_{23}c_{13}\end{array}\right), (6)

where, we are employing the notation sij=sinθijs_{ij}=\sin\theta_{ij} and cij=cosθijc_{ij}=\cos\theta_{ij}. The component R(θij,δij)R(\theta_{ij},\delta_{ij}) is a 3×33\times 3 matrix that represents a rotation in the ijij-plane by an angle θij\theta_{ij}, with their respective CP phases. The non-trivial entries can be explicitly written as [8]

R(θij)=(cijsijsijcij);R(θij,δij)=(cijsijeiδijsijeiδijcij).\displaystyle R(\theta_{ij})=\begin{pmatrix}c_{ij}&s_{ij}\\ -s_{ij}&c_{ij}\end{pmatrix};\hskip 17.07182ptR(\theta_{ij},\delta_{ij})=\begin{pmatrix}c_{ij}&s_{ij}e^{-i\delta_{ij}}\\ -s_{ij}e^{i\delta_{ij}}&c_{ij}\end{pmatrix}. (7)

This matrix is essential to comprehend how neutrino mixing works. One important feature of the UU matrix is that it is unitary (UU=IU^{\dagger}U=I), and this leads to the conservation of the transition probability, which will be employed later.

Given the hierarchical characteristic of the neutrino parameters, there are a number of approximations that are commonly used. Here we present in a concise way the main ones.

II.1 Large mass scale (atmospheric)

After adding the brief comment about the vacuum oscillation probability, we are able to explore some limits that can be used in the context of the advancement of Eq. (2). Now, for L=0L=0 we obviously get Pαβ=δαβP_{\alpha\beta}=\delta_{\alpha\beta}. As LL increases, the effects of the larger Δmij2\Delta m^{2}_{ij} will show up first, and since the mass differences are hierarchical, we may start from the regime where:

Δm312L4E=Δm322L4E1;Δm212L4E=0.\frac{\Delta m^{2}_{31}L}{4E}=\frac{\Delta m^{2}_{32}L}{4E}\gtrsim 1~{}~{}~{};~{}~{}~{}\frac{\Delta m^{2}_{21}L}{4E}=0.

Then, for a 3-neutrino scenario, the probability is

Pαβ=δαβ\displaystyle P_{\alpha\beta}=\delta_{\alpha\beta} \displaystyle- 4(j<3[Uα3Uβ3UαjUβj])sin2(ΔM2L4E)\displaystyle 4\left(\sum_{j<3}\Re\left[U^{*}_{\alpha 3}U_{\beta 3}U_{\alpha j}U^{*}_{\beta j}\right]\right)\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right) (8)
+\displaystyle+ 2(j<3[Uα3Uβ3UαjUβj])sin(ΔM2L2E),\displaystyle 2\left(\sum_{j<3}\Im\left[U^{*}_{\alpha 3}U_{\beta 3}U_{\alpha j}U^{*}_{\beta j}\right]\right)\sin\left(\frac{\Delta M^{2}L}{2E}\right),

where ΔM2=Δm312=Δm322\Delta M^{2}=\Delta m^{2}_{31}=\Delta m^{2}_{32}.

Reactor experiments:

We now want to employ Eq. (8) in experimental scenarios. For instance, in a nuclear reactor, the beta decay of fission products is an abundant source of  ν¯e\overline{\nu}_{e}. Thus, reactor experiments [9], e.g. Double-Chooz, Daya-Bay and Reno, usually study the survival probability of electron anti-neutrinos

Pee\displaystyle P_{ee} =\displaystyle= 14(|Ue3|2|Ue1|2+|Ue3|2|Ue2|2)sin2(ΔM2L4E)\displaystyle 1-4\left(|U_{e3}|^{2}|U_{e1}|^{2}+|U_{e3}|^{2}|U_{e2}|^{2}\right)\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right)
=\displaystyle= 14|Ue3|2(1|Ue3|2)sin2(ΔM2L4E),\displaystyle 1-4|U_{e3}|^{2}\left(1-|U_{e3}|^{2}\right)\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right),

which, replacing the mixing angles given by Eq. (6), it is possible to achieve the familiar form

Pee=1sin2(2θ13)sin2(ΔM2L4E).P_{ee}=1-\sin^{2}(2\theta_{13})\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right). (9)

There are no reactor experiments measuring a non-electronic appearance in an electronic neutrino flux, but the probabilities can be calculated by the same procedure. However, there is a more elegant way to arrive at these probabilities. First, consider the time evolution equation

iddt(νeνμντ)=H(νeνμντ),i\frac{d}{dt}\left(\begin{array}[]{c}\nu_{e}\\ \nu_{\mu}\\ \nu_{\tau}\end{array}\right)=H\left(\begin{array}[]{c}\nu_{e}\\ \nu_{\mu}\\ \nu_{\tau}\end{array}\right),

which in mass basis, can be written as

iddt(ν1ν2ν3)=(m122E000m222E000m322E)(ν1ν2ν3).i\frac{d}{dt}\left(\begin{array}[]{c}\nu_{1}\\ \nu_{2}\\ \nu_{3}\end{array}\right)=\left(\begin{array}[]{ccc}\frac{m_{1}^{2}}{2E}&0&0\\ 0&\frac{m_{2}^{2}}{2E}&0\\ 0&0&\frac{m_{3}^{2}}{2E}\\ \end{array}\right)\left(\begin{array}[]{c}\nu_{1}\\ \nu_{2}\\ \nu_{3}\end{array}\right).

To return to flavour basis, we observe that the mixing between mass eingenstates νi\nu_{i} and flavour eigenstates να\nu_{\alpha} is given by να=Uνi\nu_{\alpha}=U\nu_{i}. Also, we can decompose the UU matrix into its rotational components, and following the notation defined in Eq. (6), we adopt the prescription UijR(θij)U_{ij}\equiv R(\theta_{ij}), resulting in the following compact parametrization:

iddt(νeνμντ)=U23U13U12(m122E000m222E000m322E)U12U13U23(νeνμντ).i\frac{d}{dt}\left(\begin{array}[]{c}\nu_{e}\\ \nu_{\mu}\\ \nu_{\tau}\end{array}\right)=U_{23}U_{13}U_{12}\left(\begin{array}[]{ccc}\frac{m_{1}^{2}}{2E}&0&0\\ 0&\frac{m_{2}^{2}}{2E}&0\\ 0&0&\frac{m_{3}^{2}}{2E}\\ \end{array}\right)U_{12}^{\dagger}U_{13}^{\dagger}U_{23}^{\dagger}\left(\begin{array}[]{c}\nu_{e}\\ \nu_{\mu}\\ \nu_{\tau}\end{array}\right).

When Δm212=m22m12=0\Delta m^{2}_{21}=m_{2}^{2}-m_{1}^{2}=0, the U12U_{12} and U12U_{12}^{\dagger} are absorbed as internal rotations that cancel out. Then, one can define the following basis

(νμντ)=U23(νμντ),\left(\begin{array}[]{c}\nu_{\mu}^{\prime}\\ \nu_{\tau}^{\prime}\end{array}\right)=U_{23}^{\dagger}\left(\begin{array}[]{c}\nu_{\mu}\\ \nu_{\tau}\end{array}\right), (10)

and rewrite the evolution equation as

iddt(νeνμντ)=U13(m22E000m22E000M22E)U13(νeνμντ),i\frac{d}{dt}\left(\begin{array}[]{c}\nu_{e}\\ \nu_{\mu}^{\prime}\\ \nu_{\tau}^{\prime}\end{array}\right)=U_{13}\left(\begin{array}[]{ccc}\frac{m^{2}}{2E}&0&0\\ 0&\frac{m^{2}}{2E}&0\\ 0&0&\frac{M^{2}}{2E}\\ \end{array}\right)U_{13}^{\dagger}\left(\begin{array}[]{c}\nu_{e}\\ \nu_{\mu}^{\prime}\\ \nu_{\tau}^{\prime}\end{array}\right),

where m=m1=m2m=m_{1}=m_{2} and M=m3M=m_{3}. The second family decouples since U13U_{13} and U13U_{13}^{\dagger} do not affect it, so the evolution equation reduces to

iddt(νeντ)=U13(m22E00M22E)U13(νeντ).i\frac{d}{dt}\left(\begin{array}[]{c}\nu_{e}\\ \nu_{\tau}^{\prime}\end{array}\right)=U_{13}\left(\begin{array}[]{ccc}\frac{m^{2}}{2E}&0\\ 0&\frac{M^{2}}{2E}\\ \end{array}\right)U_{13}^{\dagger}\left(\begin{array}[]{c}\nu_{e}\\ \nu_{\tau}^{\prime}\end{array}\right).

The survival probability PeeP_{ee} was already evaluated in Eq (9), and we know that the sum of these two probabilities (survival and oscillation to τ\tau^{\prime}) must add to one, so

Peτ=sin2(2θ13)sin2(ΔM2L4E),P_{e\tau^{\prime}}=\sin^{2}(2\theta_{13})\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right), (11)

and since by Eq. 10 we have νμ|ντ=s23\braket{\nu_{\mu}}{\nu_{\tau}^{\prime}}=s_{23} and ντ|ντ=c23\braket{\nu_{\tau}}{\nu_{\tau}^{\prime}}=c_{23}, we can write:

Peμ=s232sin2(2θ13)sin2(ΔM2L4E);Peτ=c232sin2(2θ13)sin2(ΔM2L4E).P_{e\mu}=s^{2}_{23}\sin^{2}(2\theta_{13})\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right)~{}~{};~{}~{}~{}P_{e\tau}=c^{2}_{23}\sin^{2}(2\theta_{13})\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right).

To sum up, the oscillation of electron anti-neutrinos driven by the large mass scale occurs to an admixture of muon and tau neutrinos, with weights given by θ23\theta_{23}.

Accelerator experiments:

Various accelerator experiments [10], such as Minos [11] and T2K [12], produce beams of νμ\nu_{\mu} or ν¯μ\overline{\nu}_{\mu} by accelerating and colliding protons into a target. Likewise, cosmic rays (usually protons) interact with other nuclei in the atmosphere, mostly producing muonic (anti-)neutrinos [13], which are detected by experiments such as IceCube [14] and Super-Kamiokande [15, 16]. In these contexts, the survival probability PμμP_{\mu\mu} is useful and can be derived from Eq. (8):

Pμμ\displaystyle P_{\mu\mu} =\displaystyle= 14|Uμ3|2(1|Uμ3|2)sin2(ΔM2L4E)\displaystyle 1-4|U_{\mu 3}|^{2}\left(1-|U_{\mu 3}|^{2}\right)\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right)
=\displaystyle= 14c132s232(1c132s232)sin2(ΔM2L4E)\displaystyle 1-4c^{2}_{13}s^{2}_{23}(1-c^{2}_{13}s^{2}_{23})\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right)
=\displaystyle= 1(s232sin2(2θ13)+sin2(2θ23)c134)sin2(ΔM2L4E).\displaystyle 1-(s^{2}_{23}\sin^{2}(2\theta_{13})+\sin^{2}(2\theta_{23})c^{4}_{13})\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right). (12)

Using the unitarity property of the PMNS matrix, the rows (or columns) of the matrix must satisfy the orthogonality conditions

kUμkUek=0,\sum_{k}U_{\mu k}U^{*}_{ek}=0,

which can be replaced in Eq. (II.1). This makes it possible to rewrite the appearance probabilities as

Pμe\displaystyle P_{\mu e} =\displaystyle= 4|Uμ3|2|Ue3|2sin2(ΔM2L4E)=s232sin2(2θ13)sin2(ΔM2L4E),\displaystyle 4|U_{\mu 3}|^{2}|U_{e3}|^{2}\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right)=s^{2}_{23}\sin^{2}(2\theta_{13})\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right),
Pμτ\displaystyle P_{\mu\tau} =\displaystyle= 4|Uμ3|2|Uτ3|2sin2(ΔM2L4E)=c134sin2(2θ23)sin2(ΔM2L4E).\displaystyle 4|U_{\mu 3}|^{2}|U_{\tau 3}|^{2}\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right)=c^{4}_{13}\sin^{2}(2\theta_{23})\sin^{2}\left(\frac{\Delta M^{2}L}{4E}\right).

which are related to the second and third term in Eq. (12). So the main oscillation channel of atmospheric and accelerator muon neutrinos is to tau neutrinos [17], with a small production of electron anti-neutrinos driven by the size of θ13\theta_{13}.

II.2 Small mass scale (solar)

The next approximation to be taken into account is when the solar mass scale does not vanish, but have a small value. Indeed, this limit can be defined through the relation

Δm212L4E1.\frac{\Delta m^{2}_{21}L}{4E}\gtrsim 1.

In this regime, we can write the probability of survival of electron neutrinos as a simple expansion of Eq. (2), where the term with the solar scale is fully expressed as

Pee=14|Ue2|2|Ue1|2sin2(Δm212L4E)+f,P_{ee}=1-4|U_{e2}|^{2}|U_{e1}|^{2}\sin^{2}\left(\frac{\Delta m^{2}_{21}L}{4E}\right)+f,

while the large mass scale contributions are grouped together in the following parameter:

f=4|Ue3|2[|Ue1|2sin2(Δm312L4E)+|Ue2|2sin2(Δm322L4E)].f=-4|U_{e3}|^{2}\left[|U_{e1}|^{2}\sin^{2}\left(\frac{\Delta m^{2}_{31}L}{4E}\right)+|U_{e2}|^{2}\sin^{2}\left(\frac{\Delta m^{2}_{32}L}{4E}\right)\right].

In this small mass scale regime, we can average out the oscillation driven by the larger terms Δm312L/4E\Delta m^{2}_{31}L/{4E} and Δm322L/4E\Delta m^{2}_{32}L/4E, resulting in

f2|Ue3|2(|Ue1|2+|Ue2|2)=2|Ue3|2(1|Ue3|2),f\sim-2|U_{e3}|^{2}\left(|U_{e1}|^{2}+|U_{e2}|^{2}\right)=-2|U_{e3}|^{2}\left(1-|U_{e3}|^{2}\right),

which, by replacing the explicit values of the UU matrix elements with their corresponding mixing angles, makes it possible to derive, after some algebraic manipulation, the electron neutrino survival probability:

Pee=c134Pee(2fam)+s134,P_{ee}=c^{4}_{13}P_{ee}^{(2fam)}+s^{4}_{13},

where Pee(2fam)P_{ee}^{(2fam)} is the survival probability in the limit θ130\theta_{13}\rightarrow 0:

Pee(2fam)=1sin2(2θ12)sin2(Δm212L4E).P_{ee}^{(2fam)}=1-\sin^{2}(2\theta_{12})\sin^{2}\left(\frac{\Delta m^{2}_{21}L}{4E}\right).

The only approximation used on this formula regards the mass hierarchy. If we also want to disregard θ13\theta_{13}, then the survival probability trivially reduces to the 2-family one. This is the expression used for understanding KamLAND results [18].

III Matter effects

Although very useful to describe the oscillation of terrestrial neutrinos, astrophysical neutrinos are usually created in dense environments, such as the cores of stars or supernovae. In these conditions, matter effects significantly impact their oscillation behavior, as interactions with particles in the medium modify the effective mixing angles and mass differences. But even for terrestrial neutrinos, considering matter effects on flavor conversion probabilities are crucial when precision predictions are needed, such as in the case of determining the mass ordering of the neutrino families.

The main effect of matter effects can be analyzed through a modification on the mass and mixing matrix parameters. If neutrinos travels through a constant matter environment, the formulas derived in previous sections could be replaced by the same expressions, but with the new values for mass and mixing angles calculated in matter. Finding expressions for these parameters is what we present in the following section. However, for a varying matter density, new phenomena, like resonant flavor conversion such as the MSW effect [19, 20] have to be considered, resulting in very different conversion probabilities compared to those in a vacuum. We will not focus on conversion probabilities in varying density environments in the present work.

The vacuum Hamiltonian (HvacH_{\text{vac}}) is modified to the matter Hamiltonian (HmatH_{mat}), which effectively accounts for the following displacement

Hmat=H+VCC,H_{mat}=H+V_{CC}, (13)

where the matrix VCCV_{CC} encompass the matter potential effect.

At first we present some generic expressions for two families, which can be used as an approximation in some specific scenarios for three families, presented in afterward. In what follows, unless explicitly stated otherwise, we use for the oscillation parameters the values:

Δm212=8 105eV2,Δm312=2.5 103eV2;\displaystyle\Delta m^{2}_{21}=8\,10^{-5}~{}\textrm{eV}^{2},\quad\Delta m^{2}_{31}=2.5\,10^{-3}~{}\textrm{eV}^{2};
θ12=0.59,θ13=0.148,θ23=0.738;\displaystyle\theta_{12}=0.59,\quad\theta_{13}=0.148,\quad\theta_{23}=0.738;
Eν=10MeV.\displaystyle E_{\nu}=10\,\textrm{MeV}.

III.1 Two families

Analyzing the neutrino oscillation problem within the framework of only two families yields significant benefits, as it envelopes the fundamental aspects of the mechanism while maintaining a simplified structure. Given the mixing matrix between mass eingenstates νi\nu_{i} and flavour eigenstates να\nu_{\alpha} already disscused and the rotation matrix defined in Eq (7), the Hamiltonian in flavour basis assuming that the first state in flavour base is electronic can be written as:

Hmat\displaystyle H_{mat} =\displaystyle= m12+m224E+U(Δm24Eν00+Δm2124Eν)U+(VCC000)\displaystyle\frac{m_{1}^{2}+m_{2}^{2}}{4E}+U\left(\begin{array}[]{cc}-\frac{\Delta m^{2}}{4E_{\nu}}&0\\ 0&+\frac{\Delta m^{2}_{21}}{4E_{\nu}}\\ \end{array}\right)U^{\dagger}+\left(\begin{array}[]{cc}V_{CC}&0\\ 0&0\end{array}\right) (18)
=\displaystyle= m12+m224E+VCC2+(Δm24Eνc2θ+VCC2Δm24Eνs2θΔm24Eνs2θ+Δm24Eνc2θVCC2),\displaystyle\frac{m_{1}^{2}+m_{2}^{2}}{4E}+\frac{V_{CC}}{2}+\left(\begin{array}[]{cc}-\frac{\Delta m^{2}}{4E_{\nu}}c_{2\theta}+\frac{V_{CC}}{2}&\frac{\Delta m^{2}}{4E_{\nu}}s_{2\theta}\\ \frac{\Delta m^{2}}{4E_{\nu}}s_{2\theta}&+\frac{\Delta m^{2}}{4E_{\nu}}c_{2\theta}-\frac{V_{CC}}{2}\\ \end{array}\right), (21)

where Δm2=Δm212\Delta m^{2}=\Delta m_{21}^{2}. In addition, the notation c2θ=cos2θc_{2\theta}=\cos 2\theta and s2θ=sin2θs_{2\theta}=\sin 2\theta was employed. The eigenvalues of the non-diagonal matrix can be calculated through the usual diagonalization process. Furthermore, the terms proportional to identity in Eq. (21) should be added to the full eigenvalues:

λ1,2=m12+m224E+VCC2±(Δm24Eνc2θVCC2)2+(Δm24Eνs2θ)2.\lambda_{1,2}=\frac{m_{1}^{2}+m_{2}^{2}}{4E}+\frac{V_{CC}}{2}\pm\sqrt{\left(\frac{\Delta m^{2}}{4E_{\nu}}c_{2\theta}-\frac{V_{CC}}{2}\right)^{2}+\left(\frac{\Delta m^{2}}{4E_{\nu}}s_{2\theta}\right)^{2}}. (22)

We have to diagonalize this matrix, i.e. find the modifications in the mixing angles in matter that leads to:

U~HU~=diag(λ1,λ2),\tilde{U^{\dagger}}H\tilde{U}=\textrm{diag}(\lambda_{1},\lambda_{2}), (23)

where U~\tilde{U} has the same structure as in Eq. (6), but with angles modified by matter interactions. The mixing angle in matter θ~\tilde{\theta} that reproduces HmH_{m} is

tan(2θ~)=Δm24Eνs2θΔm24Eνc2θVCC2.\tan(2\tilde{\theta})=\frac{\frac{\Delta m^{2}}{4E_{\nu}}s_{2\theta}}{\frac{\Delta m^{2}}{4E_{\nu}}c_{2\theta}-\frac{V_{CC}}{2}}. (24)

It is possible to expand Eq. 22 in ρ\rho for particular limits, in order to achieve analytical approximations for the mixing angle and eigenvalues.

Low densities: VCCΔm24Eνc2θ,Δm24Eνs2θV_{CC}\ll\frac{\Delta m^{2}}{4E_{\nu}}c_{2\theta},\frac{\Delta m^{2}}{4E_{\nu}}s_{2\theta}

In this regime, Eq. (22) can be rewritten as

λ1,2m12+m224E+VCC2±[Δm24Eν12VCCc2θ],\lambda_{1,2}\sim\frac{m_{1}^{2}+m_{2}^{2}}{4E}+\frac{V_{CC}}{2}\pm\left[\frac{\Delta m^{2}}{4E_{\nu}}-\frac{1}{2}V_{CC}c_{2\theta}\right],

which leads to the following analytical expressions for the eigenvalues

λ1=m122Eν+cθ2VCC;λ2=m222Eν+sθ2VCC.\lambda_{1}=\frac{m_{1}^{2}}{2E_{\nu}}+c_{\theta}^{2}V_{CC}~{}~{}~{};~{}~{}~{}\lambda_{2}=\frac{m_{2}^{2}}{2E_{\nu}}+s_{\theta}^{2}V_{CC}.

High densities: VCCΔm24Eνs2θ,Δm24Eνc2θV_{CC}\gg\frac{\Delta m^{2}}{4E_{\nu}}s_{2\theta},\frac{\Delta m^{2}}{4E_{\nu}}c_{2\theta}

The same prescription can be applied for the high matter density limit, where the eigenvalues are expanded in terms of ρ\rho leading to:

λ1,2m12+m224E+VCC2±[VCC2Δm24Eνc2θ],\lambda_{1,2}\sim\frac{m_{1}^{2}+m_{2}^{2}}{4E}+\frac{V_{CC}}{2}\pm\left[\frac{V_{CC}}{2}-\frac{\Delta m^{2}}{4E_{\nu}}c_{2\theta}\right],

which, once again, according to the ordering defined for the eigenvalues before, leads to analytical approximations:

λ1=m222EνΔm22Eνcθ2.;λ2=m122Eν+Δm22Eνsθ2+VCC.\lambda_{1}=\frac{m_{2}^{2}}{2E_{\nu}}-\frac{\Delta m^{2}}{2E_{\nu}}c_{\theta}^{2}.~{}~{}~{};~{}~{}~{}\lambda_{2}=\frac{m_{1}^{2}}{2E_{\nu}}+\frac{\Delta m^{2}}{2E_{\nu}}s_{\theta}^{2}+V_{CC}.

For a constant neutrino energy and varying density, we present in Fig. (1) the eigenstates and mixing angle, together with the asymptotic behavior for low and high densities.

Refer to caption
Figure 1: Exact expressions (solid lines) and approximations (dotted) considering low and high densities for mixing angle (left panel) and eigenvalues (right panel).

The behavior that is worth pointing out is that for high densities the first family eigenvalue equals the matter potential, and the mixing angle tends to π\pi, which is a characteristic that can be found when expanding the analysis to 3 neutrino families in the next section.

III.2 Three neutrino familes

In spite of the two neutrino families analysis having a well-defined and simplified description, we will work on a more realistic case, which considers the neutrinos three families. To simplify our analisys we will set δCP=0\delta_{CP}=0 in what follows. For this specific case, the explicit form of the Hamiltonian in the flavor basis can be written as

H=U23U13U12(m122Eν000m222Eν000m322Eν)U12U13U23+(VCC00000000).H=U_{23}U_{13}U_{12}\left(\begin{array}[]{ccc}\frac{m^{2}_{1}}{2E_{\nu}}&0&0\\ 0&\frac{m^{2}_{2}}{2E_{\nu}}&0\\ 0&0&\frac{m^{2}_{3}}{2E_{\nu}}\end{array}\right)U_{12}^{\dagger}U_{13}^{\dagger}U_{23}^{\dagger}+\left(\begin{array}[]{ccc}V_{CC}&0&0\\ 0&0&0\\ 0&0&0\end{array}\right). (25)

The Hamiltonian in Eq. (25) can be analyzed using the same approach as in Eq. (21), previously employed for the two-family regime. However, the key distinction in the three-family case lies in the presence of an additional eigenvalue, arising from the increased dimensionality of the system. The analysis will explore the behavior of the system in the low, intermediate, and high-density limits.

Low densities: VCCsin(2θ13)<Δm3124EνV_{CC}\sin(2\theta_{13})<\frac{\Delta m^{2}_{31}}{4E_{\nu}}, Δm2124Eν\frac{\Delta m^{2}_{21}}{4E_{\nu}}

Since VCCV_{CC} only affects H11H_{11}, the mixing angle θ23\theta_{23} can be rotated out redefining να=U23να\nu_{\alpha}^{\prime}=U_{23}^{\dagger}\nu_{\alpha}. If we also rotate out θ13\theta_{13}, defining a να′′=U13να=U13U23να\nu_{\alpha}^{\prime\prime}=U_{13}^{\dagger}\nu_{\alpha}^{\prime}=U_{13}^{\dagger}U_{23}^{\dagger}\nu_{\alpha} we get:

H′′=m12+m224Eν+(δ21c2θ12δ21s2θ120δ21s2θ12δ21c2θ12000δ31+δ32)+VCC(c1320s13c13000s13c130s132).\displaystyle H^{{}^{\prime\prime}}=\frac{m_{1}^{2}+m_{2}^{2}}{4E_{\nu}}+\left(\begin{array}[]{ccc}-\delta_{21}c_{2\theta_{12}}&\delta_{21}s_{2\theta_{12}}&0\\ \delta_{21}s_{2\theta_{12}}&\delta_{21}c_{2\theta_{12}}&0\\ 0&0&\delta_{31}+\delta_{32}\end{array}\right)+V_{CC}\left(\begin{array}[]{ccc}c^{2}_{13}&0&s_{13}c_{13}\\ 0&0&0\\ s_{13}c_{13}&0&s^{2}_{13}\end{array}\right). (32)

where δji=Δmji2/4Eν\delta_{ji}=\Delta m^{2}_{ji}/4E_{\nu}. For low densities we disregard the non-diagonal entry in the third term in right side of Eq. 32, decoupling our system into 2+1. It straightforward to obtain the third eigenvalue, related to the decoupled family:

λ3\displaystyle\lambda_{3} =\displaystyle= m322Eν+VCCs132.\displaystyle\frac{m_{3}^{2}}{2E_{\nu}}+V_{CC}s^{2}_{13}. (33)

The remaining problem is solved by the diagonalization process of the non-diagonal subspace, which results in

λ1,2\displaystyle\lambda_{1,2} =\displaystyle= m12+m224E+VCCc1322±(δ21c2θ12VCCc1322)2+(δ21s2θ12)2.\displaystyle\frac{m_{1}^{2}+m_{2}^{2}}{4E}+\frac{V_{CC}c^{2}_{13}}{2}\pm\sqrt{\left(\delta_{21}c_{2\theta_{12}}-\frac{V_{CC}c^{2}_{13}}{2}\right)^{2}+\left(\delta_{21}s_{2\theta_{12}}\right)^{2}}. (34)

The reduced 2-dimensional subsystem has an analytical solution for the mixing angle, as already shown in Eq. (23), which gives

tan(2θ~12)=Δm24Eνs2θ12Δm24Eνc2θ12VCC2c132.\tan(2\tilde{\theta}_{12})=\frac{\frac{\Delta m^{2}}{4E_{\nu}}s_{2\theta_{12}}}{\frac{\Delta m^{2}}{4E_{\nu}}c_{2\theta_{12}}-\frac{V_{CC}}{2}c_{13}^{2}}. (35)

This modification leads to the following change in the mixing matrix

να′′=U~12νiU=U23U13U~12.\nu_{\alpha}^{\prime\prime}=\tilde{U}_{12}\nu_{i}~{}~{}~{}\rightarrow~{}~{}~{}U=U_{23}U_{13}\tilde{U}_{12}.

In Fig. 2 we present θ~12\tilde{\theta}_{12} and eigenvalues as a function of density. The analytical expressions are reliable up to ρ103\rho\sim 10^{3} g/cm3.

Refer to caption
Figure 2: Mixing angle θ12\theta_{12} (left panel) and eigenvalues λ\lambda’s (right panel), considering approximation for low densities (dotted) and numerical result (continuum line). The approximations are reliable up to ρ103\rho\sim 10^{3} g/cm3, marked as a vertical line.

One interesting feature is that λ1\lambda_{1} provides a good approximation in the full range. The asymptotic value for large densities is

λ1\displaystyle\lambda_{1} =\displaystyle= m12+m224E+VCCc1322VCCc132214δ21c2θ121VCCc132+(2VCCc132δ21)2\displaystyle\frac{m_{1}^{2}+m_{2}^{2}}{4E}+\frac{V_{CC}c^{2}_{13}}{2}-\frac{V_{CC}c^{2}_{13}}{2}\sqrt{1-4\delta_{21}c_{2\theta_{12}}\frac{1}{V_{CC}c^{2}_{13}}+\left(\frac{2}{V_{CC}c^{2}_{13}}\delta_{21}\right)^{2}}
\displaystyle\sim m122Es122+m222Ec122.\displaystyle\frac{m_{1}^{2}}{2E}s^{2}_{12}+\frac{m_{2}^{2}}{2E}c^{2}_{12}.

Since θ13\theta_{13} is small, this approximation should still be valid in the first resonance, θ~12=π/4\tilde{\theta}_{12}=\pi/4.

intermediate to high densities: VCCsin(2θ13)Δm3124EνΔm2124EνV_{CC}\sin(2\theta_{13})\gtrsim\frac{\Delta m^{2}_{31}}{4E_{\nu}}\gg\frac{\Delta m^{2}_{21}}{4E_{\nu}}

If we take Δm212=0\Delta m^{2}_{21}=0, we have:

H=U23U13U12(m122Eν000m122Eν000m322Eν)U12U13U23+(VCC00000000).H=U_{23}U_{13}U_{12}\left(\begin{array}[]{ccc}\frac{m^{2}_{1}}{2E_{\nu}}&0&0\\ 0&\frac{m^{2}_{1}}{2E_{\nu}}&0\\ 0&0&\frac{m^{2}_{3}}{2E_{\nu}}\end{array}\right)U_{12}^{\dagger}U_{13}^{\dagger}U_{23}^{\dagger}+\left(\begin{array}[]{ccc}V_{CC}&0&0\\ 0&0&0\\ 0&0&0\end{array}\right).

Now, besides absorbing U23U_{23} in ν\nu^{\prime}, we can perform the multiplication with U12U_{12}, arriving at:

H=U13(m122Eν000m122Eν000m322Eν)U13+(VCC00000000).H^{\prime}=U_{13}\left(\begin{array}[]{ccc}\frac{m^{2}_{1}}{2E_{\nu}}&0&0\\ 0&\frac{m^{2}_{1}}{2E_{\nu}}&0\\ 0&0&\frac{m^{2}_{3}}{2E_{\nu}}\end{array}\right)U_{13}^{\dagger}+\left(\begin{array}[]{ccc}V_{CC}&0&0\\ 0&0&0\\ 0&0&0\end{array}\right).

Now is the second family that gets decoupled, and we can use the expressions for the 2-families. In this case, the eigenvalues and mixing angles can be respectively written as

λ2,3=m12+m324E+VCC2±(Δm3124Eνc2θ13VCC2)2+(Δm3124Eνs2θ13)2,\lambda_{2,3}=\frac{m_{1}^{2}+m_{3}^{2}}{4E}+\frac{V_{CC}}{2}\pm\sqrt{\left(\frac{\Delta m_{31}^{2}}{4E_{\nu}}c_{2\theta_{13}}-\frac{V_{CC}}{2}\right)^{2}+\left(\frac{\Delta m_{31}^{2}}{4E_{\nu}}s_{2\theta_{13}}\right)^{2}}, (36)
tan(2θ~13)=Δm3124Eνs2θ13Δm3124Eνc2θ13VCC2.\tan(2\tilde{\theta}_{13})=\frac{\frac{\Delta m_{31}^{2}}{4E_{\nu}}s_{2\theta_{13}}}{\frac{\Delta m_{31}^{2}}{4E_{\nu}}c_{2\theta_{13}}-\frac{V_{CC}}{2}}. (37)

It is interesting to note that the asymptotic value of λ3\lambda_{3} for low energies agrees with Eq. (33) at first order in ρ\rho.

Refer to caption
Figure 3: Mixing angle θ13\theta_{13} (left panel) and eigenvalues λ\lambda’s (right panel), considering approximation for high densities (dotted) and numerical result (continuum line). The approximations are reliable starting at ρ103\rho\sim 10^{3} g/cm3, marked as a vertical line.

In summary, we have so far the following approximations:

  • λ1\lambda_{1} (Eq. (34)) and λ3\lambda_{3} (Eq. (36)) valid for all ranges of ρ\rho.

  • θ13\theta_{13} (Eq. (37)) also valid for all ranges of ρ\rho.

  • λ2\lambda_{2} for low densities (around first resonance) given by Eq. (34) and a different expression for high densities (around second resonance), Eq. (36).

  • θ12\theta_{12} for low densities, given by Eq. (35).

What is lacking is a expression for θ12\theta_{12} and θ23\theta_{23} for densities larger than the second resonance. Since we have reliable expressions for the other angles, it can be calculated as follows.

For very large densities θ~13π/2\tilde{\theta}_{13}\sim\pi/2, and we can write

cosθ~13ϵ;sinθ~131ϵ22,\cos\tilde{\theta}_{13}\sim\epsilon~{}~{}~{};~{}~{}~{}\sin\tilde{\theta}_{13}\sim 1-\frac{\epsilon^{2}}{2},

and the mixing matrix can be written in powers of ϵ\epsilon as

U~=(001𝒪00)+ϵ(c12s1200s23c23)+ϵ22(001c12s23s12s230c12c23s12c230),\displaystyle\tilde{U}=\left(\begin{array}[]{ccc}0&0&1\\ \lx@intercol\hfil\mathcal{O}\hfil\lx@intercol&\begin{array}[]{c}0\\ 0\end{array}\end{array}\right)+\epsilon\left(\begin{array}[]{ccc}c_{12}&s_{12}&0\\ \lx@intercol\hfil 0\hfil\lx@intercol&\begin{array}[]{c}s_{23}\\ c_{23}\end{array}\end{array}\right)+\frac{\epsilon^{2}}{2}\left(\begin{array}[]{ccc}0&0&-1\\ c_{12}s_{23}&s_{12}s_{23}&0\\ c_{12}c_{23}&s_{12}c_{23}&0\end{array}\right), (49)

where 𝒪\mathcal{O} can be written as a mixing matrix in 2 dimensions through a new angle α\alpha:

𝒪=(s12c23c12s23c12c23s12s23s12s23c12c23c12s23s12c23)=(cosαsinαsinαcosα).\mathcal{O}=\left(\begin{array}[]{cc}-s_{12}c_{23}-c_{12}s_{23}&c_{12}c_{23}-s_{12}s_{23}\\ s_{12}s_{23}-c_{12}c_{23}&-c_{12}s_{23}-s_{12}c_{23}\end{array}\right)=\left(\begin{array}[]{cc}\cos\alpha&\sin\alpha\\ -\sin\alpha&\cos\alpha\end{array}\right).

Using Eq.(25) and the new parametrization in Eq.(49), the following relation

U~diag(λ~1,λ~2,λ~3)U~=Udiag(λ1,λ2,λ3)U+VCC.\tilde{U}\textrm{diag}(\tilde{\lambda}_{1},\tilde{\lambda}_{2},\tilde{\lambda}_{3})\tilde{U}^{\dagger}=U\textrm{diag}(\lambda_{1},\lambda_{2},\lambda_{3})U^{\dagger}+V_{CC}.

can be examined. To first order in ϵ\epsilon and considering λ~3λ~1,2\tilde{\lambda}_{3}\gg\tilde{\lambda}_{1,2}, the left side becomes:

λ~1+λ~22+(Δ31+Δ322ϵλ~3s23ϵλ~3c23ϵλ~3s23ϵλ~3c23h2×2),\frac{\tilde{\lambda}_{1}+\tilde{\lambda}_{2}}{2}+\left(\begin{array}[]{ccc}\frac{\Delta_{31}+\Delta_{32}}{2}&\epsilon\tilde{\lambda}_{3}s_{23}&\epsilon\tilde{\lambda}_{3}c_{23}\\ \begin{array}[]{c}\epsilon\tilde{\lambda}_{3}s_{23}\\ \epsilon\tilde{\lambda}_{3}c_{23}\end{array}&\lx@intercol\hfil h_{2\times 2}\hfil\lx@intercol\end{array}\right),

where

h2×2=O(Δ21/200+Δ21/2)O=Δ212(cos2αsin2αsin2αcos2α),h_{2\times 2}=O\left(\begin{array}[]{cc}-\Delta_{21}/2&0\\ 0&+\Delta_{21}/2\end{array}\right)O^{\dagger}=\frac{\Delta_{21}}{2}\left(\begin{array}[]{cc}-\cos 2\alpha&\sin 2\alpha\\ \sin 2\alpha&\cos 2\alpha\end{array}\right),

and Δji=(λ~jλ~i)/2\Delta_{ji}=(\tilde{\lambda}_{j}-\tilde{\lambda}_{i})/2. Comparing with the left side of Eq. (25), we can finally write:

tanθ~23=(Hmat)12/(Hmat)13.\tan\tilde{\theta}_{23}=(H_{mat})_{12}/(H_{mat})_{13}.

Explicitly, after some algebraic manipulation, we obtain the following expression

tanθ~23=(Δ+δcos2θ12)sin2θ13s23+2δc13sin2θ12c23(Δ+δcos2θ12)sin2θ13c232δc13sin2θ12s23.\tan\tilde{\theta}_{23}=\frac{(\Delta+\delta\cos 2\theta_{12})\sin 2\theta_{13}s_{23}+2\delta c_{13}\sin 2\theta_{12}c_{23}}{(\Delta+\delta\cos 2\theta_{12})\sin 2\theta_{13}c_{23}-2\delta c_{13}\sin 2\theta_{12}s_{23}}.
Refer to caption
Figure 4: Approximations considering high densities for θ23\theta_{23} and θ12\theta_{12}. The asymptotic analytical expressions are marked as horizontal lines.

The relation between α\alpha and mixing angles θ12\theta_{12} and θ23\theta_{23} can be written realizing that:

𝒪=(s12c23c12s23c12c23s12s23s12s23c12c23c12s23s12c23)=(s12c12c12s12)(c23s23s23c23),\mathcal{O}=\left(\begin{array}[]{cc}-s_{12}c_{23}-c_{12}s_{23}&c_{12}c_{23}-s_{12}s_{23}\\ s_{12}s_{23}-c_{12}c_{23}&-c_{12}s_{23}-s_{12}c_{23}\end{array}\right)=\left(\begin{array}[]{cc}-s_{12}&c_{12}\\ -c_{12}&-s_{12}\end{array}\right)\left(\begin{array}[]{cc}c_{23}&s_{23}\\ -s_{23}&c_{23}\end{array}\right),

so α\alpha can be related to a rotation of θ23\theta_{23} followed by a rotation of θ12+π/2\theta_{12}+\pi/2. Then

θ~12=απ2θ~23,\tilde{\theta}_{12}=\alpha-\frac{\pi}{2}-\tilde{\theta}_{23},

so, after finding θ~23\tilde{\theta}_{23} it is straightforward to find the value of θ~12\tilde{\theta}_{12}. In Fig. (4) we can see that the asymptotic value agrees with the numerical calculation.

IV Conclusions

In this article, we present analytical expressions for neutrino flavor conversion probabilities and mixing angles in matter across different experimental contexts. The matter dependence of the mixing angle θ23\theta_{23} is analyzed, a result not found in the main reviews on the neutrino phenomenology. Designed as an introductory guide for students new to the field of Neutrino Physics, this text is supported by numerical programs available in the GitHub repository of our research group.

References