Raffelt - Stars As Laboratories For Fundamental Physics
Raffelt - Stars As Laboratories For Fundamental Physics
Raffelt - Stars As Laboratories For Fundamental Physics
Raffelt
Stars as Laboratories
for Fundamental Physics
The Astrophysics of Neutrinos, Axions, and Other
Weakly Interacting Particles
Preface (xiv)
Acknowledgments (xxii)
1. The Energy-Loss Argument
1.1 Introduction (1)
1.2 Equations of Stellar Structure (5)
1. Hydrostatic Equilibrium (5) ! 2. Generic Cases of Stellar Structure (7) !
3. Energy Conservation (10) ! 4. Energy Transfer (11) ! 5. Gravitational
Settling (13)
1.3 Impact of Novel Particles (14)
1. Energy Loss (14) ! 2. Application to the Sun (16) ! 3. Radiative Energy
Transfer (17) ! 4. Opacity Contribution of Arbitrary Bosons (18) ! 5. Solar
Bound on Massive Pseudoscalars (20)
1.4 General Lesson (21)
v
vi Table of Contents
7. Nonstandard Neutrinos
7.1 Neutrino Masses (251)
1. The Fermion Mass Problem (251) ! 2. Dirac and Majorana Masses (253) !
3. Kinematical Mass Bounds (254) ! 4. Neutrinoless Double-Beta Decay (257)
! 5. Cosmological Mass Bounds (258).
7.2 Neutrino Mixing and Decay (260)
1. Flavor Mixing (260) ! 2. Standard-Model Decays of Mixed Neutrinos (263)
7.3 Neutrino Electromagnetic Form Factors (267)
1. Overview (267) ! 2. Single-Photon Coupling (268) ! 3. Two-Photon
Coupling (271)
7.4 Electromagnetic Processes (272)
7.5 Limits on Neutrino Dipole Moments (275)
1. Scattering Experiments (275) ! 2. Spin-Flip Scattering in Supernovae (277)
! 3. Spin-Flip Scattering in the Early Universe (277) ! 4. Search for Radiative
Neutrino Decays (278) ! 5. Plasmon Decay in Stars (279).
Table of Contents ix
8. Neutrino Oscillations
8.1 Introduction (280)
8.2 Vacuum Oscillations (282)
1. Equation of Motion for Mixed Neutrinos (282) ! 2. Two-Flavor Oscilla-
tions (284) ! 3. Distribution of Sources and Energies (287) ! 4. Experimental
Oscillation Searches (289) ! 5. Atmospheric Neutrinos (290)
8.3 Oscillations in Media (293)
1. Dispersion Relation for Mixed Neutrinos (293) ! 2. Oscillations in Homo-
geneous Media (296) ! 3. Inhomogeneous Medium: Adiabatic Limit (297)
! 4. Inhomogeneous Medium: Analytic Results (299) ! 5. The Triangle and
the Bathtub (301) ! 6. Neutrino Oscillations without Vacuum Mixing (303)
8.4 Spin and Spin-Flavor Oscillations (304)
1. Vacuum Spin Precession (304) ! 2. Spin Precession in a Medium (306) !
3. Spin-Flavor Precession (306) ! 4. Twisting Magnetic Fields (308)
14. Axions
14.1 The Strong CP-Problem (524)
14.2 The Peccei-Quinn Mechanism (521)
1. Generic Features (526) ! 2. Axions as Nambu-Goldstone Bosons (528)
! 3. Pseudoscalar vs. Derivative Interaction (531) ! 4. The Onslaught of
Quantum Gravity (532)
14.3 Fine Points of Axion Properties (534)
1. The Most Common Axion Models (534) ! 2. Axion Mass and Coupling
to Photons (535) ! 3. Model-Dependent Axion-Fermion Coupling (536)
14.4 Astrophysical Axion Bounds (538)
14.5 Cosmological Limits (540)
Appendices
A. Units and Dimensions (580)
B. Neutrino Coupling Constants (583)
C. Numerical Neutrino Energy-Loss Rates (585)
1. Plasma Process (585) ! 2. Photoneutrino and Pair-Annihilation Pro-
cess (586) ! 3. Bremsstrahlung (588) ! 4. Total Emission Rate (589)
References (606)
Acronyms (642)
Symbols (644)
Subject Index (649)
Preface
Ever since Newton proposed that the moon on its orbit follows the same
laws of motion as an apple falling from a tree, the heavens have been
a favorite laboratory to test the fundamental laws of physics, notably
classical mechanics and Newton’s and Einstein’s theories of gravity.
This tradition carries on—the 1993 physics Nobel prize was awarded
to R. A. Hulse and J. H. Taylor for their 1974 discovery of the binary
pulsar PSR 1913+16 whose measured orbital decay they later used to
identify gravitational wave emission. However, the scope of physical
laws necessary to understand the phenomena observed in the super-
lunar sphere has expanded far beyond these traditional fields. Today,
astrophysics has become a vast playing ground for applications of the
laws of microscopic physics, in particular the properties of elementary
particles and their interactions.
This book is about how stars can be used as laboratories to probe
fundamental interactions. Apart from a few arguments relating to grav-
itational physics and the nature of space and time (Is Newton’s constant
constant? Do all relativistic particles move with the same limiting ve-
locity? Are there novel long-range interactions?), most of the discussion
focusses on the properties and nongravitational interactions of elemen-
tary particles.
There are three predominant methods for the use of stars as particle-
physics laboratories. First, stars are natural sources for photons and
neutrinos which can be detected on Earth. Neutrinos are now routinely
measured from the Sun, and have been measured once from a collapsing
star (SN 1987A). Because these particles literally travel over astronom-
ical distances before reaching the detector one can study modifications
of the measured signal which can be attributed to propagation and dis-
persion effects, including neutrino flavor oscillations or axion-photon
oscillations in intervening magnetic fields. It is well known that the
discrepancy between the calculated and measured solar neutrino spec-
tra is the most robust, yet preliminary current indication for neutrino
oscillations and thus for nonvanishing neutrino masses.
xiv
Preface xv
Second, particles from distant sources may decay, and there may
be photons or even measurable neutrinos among the decay products.
The absence of solar x- and γ-rays yields a limit on neutrino radia-
tive decays which is as “safe” as a laboratory limit, yet nine orders
of magnitude more restrictive. An even more restrictive limit obtains
from the absence of γ-rays in conjunction with the SN 1987A neutri-
nos which allows one to conclude, for example, that even ντ must obey
the cosmological limit of mν < ∼ 30 eV unless one invents new invisible
decay channels.
Third, the emission of weakly interacting particles causes a direct
energy-loss channel from the interior of stars. For neutrinos, this effect
has been routinely included in stellar evolution calculations. If new low-
mass elementary particles were to exist such as axions or other Nambu-
Goldstone bosons, or if neutrinos had novel interactions with the stellar
medium such as one mediated by a putative neutrino magnetic dipole
moment, then stars might lose energy too fast. A comparison with the
observed stellar properties allows one to derive restrictive limits on the
operation of a new energy-loss or energy-transfer mechanism and thus
to constrain the proposed novel particle interactions.
While these and related arguments as well their application and re-
sults are extensively covered here, I have not written on several topics
that might be expected to be represented in a book on the connection
between particle physics and stars. Neutron stars have been speculated
to consist of quark matter so that in principle they are a laboratory to
study a quark-gluon plasma. As I am not familiar enough with the liter-
ature on this interesting topic I refer the reader to the review by Alcock
and Olinto (1988) as well as to the more recent proceedings of two top-
ical conferences (Madsen and Haensel 1992; Vassiliadis et al. 1995).
I have also dodged some important issues in the three-way relation-
ship between cosmology, stars, and particle physics. If axions are not
the dark matter of the universe, it is likely filled with a “background
sea” of hypothetical weakly interacting massive particles (WIMPs) such
as the lightest supersymmetric particles. Moreover, there may be ex-
otic particles left over from the hot early universe such as magnetic
monopoles which are predicted to exist in the framework of typical
grand unified theories (GUTs). Some of the monopoles or WIMPs
would be captured and accumulate in the interior of stars. GUT
monopoles are predicted to catalyze nucleon decay (Rubakov-Callan-
effect), providing stars with a novel energy source. This possibility can
be constrained by analogous methods to those presented here which
limit anomalous energy losses. The resulting constraints on the pres-
xvi Preface
ence of GUT monopoles in the universe have been reviewed, for exam-
ple, in the cosmology book of Kolb and Turner (1990). Because nothing
of substance has changed, a new review did not seem warranted.
WIMPs trapped in stars would contribute to the heat transfer be-
cause their mean free path can be so large that they may be orbit-
ing almost freely in the star’s gravitational potential well, with only
occasional collisions with the background medium. Originally it was
thought that this effect could reduce the central solar temperature
enough to solve the solar neutrino problem, and to better an alleged
discrepancy between observed and predicted solar p-mode frequencies.
With the new solar neutrino data it has become clear, however, that a
reduction of the central temperature alone cannot solve the problem.
Worse, solving the “old solar neutrino problem” by the WIMP mecha-
nism now seems to cause a discrepancy with the observed solar p-mode
frequencies (Christensen-Dalsgaard 1992). In addition, a significant ef-
fect requires relatively large scattering cross sections and thus rather
contrived particle-physics models. Very restrictive direct laboratory
constraints exist for the presence of these “cosmions” in the galaxy.
Given this status I was not motivated to review the topic in detail.
The annihilation of dark-matter WIMPs captured in the Sun or
Earth produces high-energy neutrinos which are measurable in terres-
trial detectors such as Kamiokande or the future Superkamiokande,
NESTOR, DUMAND, and AMANDA Cherenkov detectors. This in-
direct approach to search for dark matter may well turn into a serious
competitor for the new generation of direct laboratory search experi-
ments that are currently being mounted. This material is extensively
covered in a forthcoming review Supersymmetric Dark Matter by Jung-
man, Kamionkowski, and Griest (1995); there is no need for me to
duplicate the effort of these experts.
The topics covered in my book revolve around the impact of low-
mass or massless particles on stars or the direct detection of this ra-
diation. The highest energies encountered are a few 100 MeV (in the
interior of a SN core) which is extremely small on the high-energy scales
of typical particle accelerator experiments. Therefore, stars as labora-
tories for fundamental physics help to push the low-energy frontier of
particle physics and as such complement the efforts of nonaccelerator
particle experiments. Their main thrust is directed at the search for
nonstandard neutrino properties, but there are other fascinating top-
ics which include the measurement of parity-violating phenomena in
atoms, the search for neutron or electron electric dipole moments which
would violate CP, the search for neutron-antineutron oscillations, the
Preface xvii
search for proton decay, or the search for particle dark matter in the
galaxy by direct and indirect detection experiments (Rich, Lloyd Owen,
and Spiro 1987). It is fascinating that the IMB and Kamiokande water
Cherenkov detectors which had been built to search for proton decay
ended up seeing supernova (SN) neutrinos instead. The Fréjus detec-
tor, instead of seeing proton decay, has set important limits on the
oscillation of atmospheric neutrinos. Kamiokande has turned into a
major solar neutrino observatory and dark-matter search experiment.
The forthcoming Superkamiokande and SNO detectors will continue
and expand these missions, may detect a future galactic SN, and may
still find proton decay.
"
"
"
scales that have been used for the purposes of particle astrophysics.
For each case the salient applications are summarized. Chapters 3−6
deal with the interaction of “radiation” (neutrinos, axions, other low-
mass bosons) with the main constituents of stellar plasmas (photons,
electrons, nucleons). In these chapters all information is pulled together
that pertains to the given interaction channel, even if it is not directly
related to the energy-loss argument. For example, in Chapter 5 the
limits on the electromagnetic coupling of pseudoscalars with photons
are summarized; the stellar energy-loss ones are the most restrictive
which justifies this arrangement.
Chapter 6 develops the topic of particle dispersion in media. The
medium-induced photon dispersion relation allows for the plasma pro-
cess γ → νν which yields the best limit on neutrino dipole moments
by virtue of the energy-loss argument applied to globular-cluster stars.
The neutrino dispersion relation is needed for the following discussion
of neutrino oscillations, establishing a link between the energy-loss ar-
gument and the dispersion arguments of the following chapters.
Dispersion and propagation effects are particularly important for
massive neutrinos with flavor mixing. To this end the phenomenology
of massive, mixed neutrinos is introduced in Chapter 7. Vacuum and
matter-induced flavor oscillations as well as magnetically induced spin
oscillations are taken up in Chapter 8. If neutrinos are in thermal equi-
librium as in a young SN core or the early universe, neutrino oscillations
require a different theoretical treatment (Chapter 9).
Chapters 10−13 are devoted to astrophysical sources where neu-
trinos have been measured, i.e. the Sun and supernovae (for the lat-
ter only the SN 1987A signal exists). Neutrino oscillations, notably
of the matter-induced variety, play a prominent role in Chapter 10
(solar neutrinos) and Chapter 11 (SN neutrinos). Radiative particle
decays, especially of neutrinos, are studied in Chapter 12 where the
Sun and SN 1987A figure prominently as sources. In Chapter 13 the
particle-physics results from SN 1987A are summarized, including the
ones related to the energy-loss and other arguments.
Chapters 14−16 give particle-specific summaries. While axions play
a big role throughout this text, only the structure of the interaction
Hamiltonian with photons, electrons, and nucleons is needed. Thus
everything said about axions applies to any pseudoscalar low-mass bo-
son for which they serve as a generic example. In Chapter 14 these
results are interpreted in terms of axion-specific models which relate
their properties to those of the neutral pion and thus establish a nearly
unique relationship between their mass and interaction strength. Often-
Preface xxi
"
All those whose lives are spent searching for truth are well
aware that the glimpses they catch of it are necessarily fleet-
ing, glittering for an instant only to make way for new and
still more dazzling insights. The scholar’s work, in marked
contrast to that of the artist, is inevitably provisional. He
knows this and rejoices in it, for the rapid obsolescence of
his books is the very proof of the progress of scholarship.
(Henri Pirenne, 1862−1935)
While writing this book I have benefitted from the help and encourage-
ment of many friends and colleagues. In particular, I need to mention
Hans-Thomas Janka, Lothar Luh, Günter Sigl, Pierre Sikivie, Thomas
Strobel, and Achim Weiss who read various parts of the manuscript,
and the referees Josh Frieman and J. Craig Wheeler who read all or
most of it. Their comments helped in no small measure to improve the
manuscript and to eliminate some errors. Hans-Thomas, in particular,
has spared no effort at educating me on the latest developments in the
area of supernova physics. Several chapters were written during a visit
at the Center for Particle Astrophysics in Berkeley—I gratefully ac-
knowledge the fine hospitality of Bernard Sadoulet and his staff. Stay-
ing for that period as a guest in Edward Janelli’s house made a huge
difference. During the entire writing process Greg Castillo provided
encouragement and a large supply of Cuban and Puerto Rican music
CDs which have helped to keep my spirits up. At the University of
Chicago Press, I am indebted to Vicki Jennings, Penelope Kaiserlian,
Stacia Kozlowski, and Eleanore Law for their expert handling of all
editorial and practical matters that had to be taken care of to trans-
form my manuscript into a book. David Schramm as the series editor
originally solicited this opus (“just expand your Physics Report a little
bit”). It hasn’t quite worked that way, but now that I’m finished I am
grateful that David persuaded me to take up this project.
xxii
Chapter 1
1.1 Introduction
More than half a century ago, Gamow and Schoenberg (1940, 1941)
ushered in the advent of particle astrophysics when they speculated that
neutrinos may play an important role in stellar evolution, particularly
in the collapse of evolved stars. Such a hypothesis was quite bold for
the time because neutrinos, which had been proposed by Pauli in 1930,
were not directly detected until 1954. That their existence was far from
being an established belief when Gamow and Schoenberg wrote their
papers is illustrated by Bethe’s (1939) complete silence about them in
his seminal paper on the solar nuclear fusion chains.
Even after the existence of neutrinos had been established they
seemed to interact only by β reactions of the sort e− + (A, Z) →
(A, Z−1) + νe or (A, Z−1) → (A, Z) + e− + ν e , the so-called URCA
reactions which Gamow and Schoenberg had in mind, or by fusion pro-
cesses like pp → de+ νe . The URCA reactions and related processes
become important only at very high temperatures or densities because
of their energy threshold. While the Sun emits two neutrinos for every
helium nucleus fused from hydrogen, the energy loss in neutrinos is only
a few percent of the total luminosity and thus plays a minor role.
1
2 Chapter 1
As a simple example for the beauty and power of the virial theo-
rem one may estimate the solar central temperature from its mass and
radius. The material is dominated by protons which have a gravita-
tional potential energy of order −GN M" mp /R" = −2.14 keV where
M" = 1.99×1033 g is the solar mass, R" = 6.96×1010 cm the solar
radius, and mp the proton mass. The average kinetic energy of a
proton is equal to 32 T (remember, kB has been set equal to unity),
yielding an approximate value for the solar internal temperature of
T = 13 2.14 keV = 0.8×107 K. This is to be compared with 1.56×107 K
found for the central temperature of a typical solar model. This exam-
ple illustrates that the basic properties of stars can be understood from
simple physical principles.
a) Normal Stars
There are two main sources of pressure relevant in stars, thermal pres-
sure and degeneracy pressure. The third possibility, radiation pressure,
never dominates except perhaps in the most massive stars. The pres-
sure provided by a species of particles is proportional to their density,
to their momentum which is reflected on an imagined piston and thus
exerts a force, and to their velocity which tells us the number of hits on
the piston per unit time. In a nondegenerate nonrelativistic medium
a typical particle velocity and momentum is proportional to T 1/2 so
that p ∝ (ρ/µ) T with ρ the mass density and µ the mean molecu-
lar weight of the medium constituents. For nonrelativistic degenerate
electrons the density is ne = p3F /3π 2 (Fermi momentum pF ), a typical
momentum is pF , and the velocity is pF /me , yielding a pressure which
is proportional to p5F or to n5/3
e and thus to ρ5/3 .
The two main pressure sources determine two generic forms of be-
havior of overall stellar models, namely normal stars such as our Sun
which is dominated by thermal pressure, and degenerate stars such as
white dwarfs which are dominated by degeneracy pressure. These two
cases follow a very different logic.
A normal star is understood most easily if one imagines how it ini-
tially forms from a dispersed but gravitationally bound gas cloud. It
continuously loses energy because photons are produced in collisions
between, say, electrons and protons. The radiation carries away energy
which must go at the expense of the total energy of the system. If it is
roughly in an equilibrium configuration, the virial theorem Eq. (1.2) in-
8 Chapter 1
b) Degenerate Stars
Everything is different for a configuration dominated by degeneracy
pressure. Above all, it has a positive heat capacity so that a loss
of energy no longer implies contraction and heating. The star actu-
ally cools. This is what happens to a brown dwarf which is a star
so small (M < 0.08 M" ) that it did not reach the critical condi-
tions to ignite hydrogen: it becomes a degenerate gas ball which slowly
“browns out.”
The relationship between radius and mass is inverted. A normal
star is geometrically larger if it has a larger mass; very crudely R ∝
M. When mass is added to a degenerate configuration it becomes
geometrically smaller as the reduced size squeezes the electron Fermi
sea into higher momentum states, providing for increased pressure to
balance the increased gravitational force. The l.h.s. of Eq. (1.1) can be
approximated as p/R where p ∝ ρ5/3 is some average pressure. Because
ρ ≈ M/R3 one finds p/R ∝ M5/3 /R6 while the r.h.s. of Eq. (1.1) is
proportional to Mρ/R2 and thus to M2 /R5 . Therefore, a degenerate
configuration is characterized by R ∝ M−1/3 .
Increasing the mass beyond a certain limit causes the radius to
shrink so much that the electrons become relativistic. Then they move
with a velocity fixed at c (or 1 in natural units), causing the pressure
to vary only as p4F or ρ4/3 . In this case adding mass no longer leads to a
sufficient pressure increase to balance for the extra weight. Beyond this
“Chandrasekhar limit,” which is about 1.4 M" for a chemical composi-
tion with Ye = 12 (number of electrons per baryon), no stable degenerate
configuration exists.
In summary, the salient features of a degenerate configuration are
the inverse mass-radius relationship R ∝ M−1/3 , the Chandrasekhar
limit, the absence of nuclear burning, and the positive specific heat
which allows the configuration to cool when it loses energy.
10 Chapter 1
c) Giant Stars
A real star can be a hybrid configuration with a degenerate core and
a nondegenerate envelope with nuclear burning at the bottom of the
envelope, i.e. the surface of the core. The core then follows the logic
of a degenerate configuration. The envelope follows the self-regulating
logic of a normal star except that it is no longer dominated by its self-
gravity but rather by the gravitational force exerted by the compact
core. Amazingly, the envelopes of such stars tend to expand to huge
dimensions. It does not seem possible to explain in a straightforward
way why stars become giants2 except that “the equations say so.” Low-
mass red giants will play a major role in Chapter 2—a further discussion
of their fascinating story is deferred until then.
4πr2 d(aT 4 )
Lr = − , (1.6)
3κρ dr
κ−1 = κ−1 −1 −1
γ + κc + κx , (1.7)
ing M; stars with M < ∼ 0.25 M" are fully convective. For M ∼ M"
>
the outer regions are radiative while the core is convective out to an ever
increasing mass fraction of the star with increasing M. A star with M
near 1 M" is very special in that it is radiative almost throughout; the
Sun is thought to have only a relatively minor convective surface layer.
Besides transporting energy, convection also moves matter and thus
affects the composition profile of a star. This is seen, for example, in
the upper panels of Fig. 2.4 where the hydrogen depletion of a solar
model (which is radiative) is a function of the local nuclear burning
rates while for the convective helium core of a horizontal-branch (HB)
star the helium depletion reaches to much larger radii than nuclear
burning. The long lifetimes of HB stars cannot be understood without
the convective supply of fuel to the nuclear furnace at the center. The
Sun, on the other hand, will complete its main-sequence evolution when
hydrogen is depleted at the center, corresponding to about a 10% global
depletion only.
The extent of convective regions can change during the course of
stellar evolution. They can leave behind composition discontinuities
which are a memory of a previous configuration. For example, on the
lower red-giant branch (RGB) the convective envelope reaches so deep
that it penetrates into the region of variable hydrogen content caused
by nuclear burning. Later, the convective envelope retreats from the ad-
vancing hydrogen-burning shell which encounters a discontinuity in the
hydrogen profile. This causes a brief “hesitation” on the RGB ascent
and thus a “bump” in the distribution of stars in the color-magnitude
diagram of globular clusters on the lower RGB. This bump has been
identified in several clusters (Figs. 2.18 and 2.19); its location is in good
agreement with theoretical expectations (Fusi Pecci et al. 1990).
Table 1.1. Initial helium abundance for solar models with axion losses.
g10 Yinitial δx Xc
0 0.274 0.00 0.362
10 0.266 0.16 0.307
15 0.256 0.32 0.292
20 0.241 0.51 0.245
25 0.224 0.65 0.151
Sect. 5.2. For the present discussion the axion losses represent some
generic energy-loss mechanism with a rate proportional to the square
of the axion-photon coupling strength gaγ .
Without exotic losses a presolar helium abundance of Yinitial = 0.274
was needed to reproduce the present-day Sun. For several values of
g10 ≡ gaγ /10−10 GeV−1 Raffelt and Dearborn found the initial helium
values given in Tab. 1.1 necessary to produce the present-day luminos-
ity. The values for δx in Tab. 1.1 are defined as in Eq. (1.15) with Lx
the axion luminosity of the (perturbed) present-day solar model which
has Lγ = L" . Also, the central hydrogen abundance Xc of the present-
day model is given. For g10 = 30, corresponding to δx ≈ 0.75, no
present-day Sun could be constructed for any value of Yinitial .
The primordial helium abundance is thought to be about 23%, and
the presolar abundance is certainly larger. Still, a value of δx less than
about 0.5 is hard to exclude on the basis of this calculation. Therefore,
the approximate solar constraint remains δx < 1 <
∼ 2 or Lx ∼ L" as found
from the analytic treatment in the previous section.
One may be able to obtain an interesting limit by considering the
oscillation frequencies of the solar pressure modes. Because of the excel-
lent agreement between standard solar models and the observed p-mode
frequencies there is little leeway for a modified solar structure and com-
position. This method has been used to constrain a hypothetical time
variation of Newton’s constant (Sect. 15.2.3).
such large interactions that they are even trapped, say, in the Sun.
Because their mean free path (mfp) is now less than the geometric di-
mension of the star, they remove energy from one region and deposit it
at an approximate distance of one mfp. This is precisely the mechanism
of radiative energy transfer: the particles now contribute to the opac-
ity. In a transition region where the mfp is on the order of the stellar
radius this mode of energy transfer couples distant regions and thus
cannot be described in the form of the differential equation (1.6). In
this case the difference between an energy-loss and an energy-transfer
mechanism is blurred.
Equation (1.6) is justified when the mfp is less than the temper-
ature scale height (d ln T /dr)−1 . For radiative transfer (κρ)−1 is an
average mfp. Therefore, Eq. (1.6) informs us that the energy flux car-
ried through a sphere of radius r is a product of a numerical factor, the
area, the photon mfp, the number density of photons, and the temper-
ature gradient. Radiative transfer is more efficient for a larger mfp, i.e.
for a more weakly interacting particle!
If a second photon existed with a coupling strength α# instead of
1
α = 137 , it would contribute more to the energy transfer for α# < α.
The observed properties of the Sun and other stars confirm that the
standard opacities certainly cannot be wrong by more than a factor
of a few. Therefore, the new photon must interact about as strongly
as the standard one to be in agreement with the properties of stars,
or it must interact so weakly that it freely escapes and the integrated
volume emission Lγ ! is less than the photon luminosity Lγ .
160ααx Ye T 6
(x = F (mx /T ), (1.21)
π mu m4e
√
where F (z) = 1 for mx + T and F (z) = (20 2)−1 z 9/2 e−z for T +
mx + me . Further, Ye is the number of electrons per baryon and mu is
the atomic mass unit. (It was used that approximately ne /ρ = Ye /mu
with the electron density ne .)
For T + mx + me the Compton absorption rate is found to be
Γx = 4πααx m2x m−4
e ne , leading to an opacity contribution of the pseudo-
Fig. 1.1. Effects of massive pseudoscalar particles on the Sun (interior tem-
perature about 1 keV). Above the dashed line they contribute to the radia-
tive energy transfer, below they escape freely and drain the Sun of energy.
The shaded area is excluded by this simple argument. (Adapted from Raffelt
and Starkman 1989.)
The Energy-Loss Argument 21
scalars of
2(2π)9/2 ααx Ye T 5/2 emx /T
κx = 1/2
45 mx m4e
= 4.4×10−3 cm2 g−1 × αx Ye m−0.5 2.5 mx /T
keV TkeV e , (1.22)
where mkeV = mx /keV and TkeV = T /keV.
The observed properties of the Sun then allow one to exclude a large
range of parameters in the mx -αx -plane. The energy-loss rate integrated
over the entire Sun must not exceed L" . Moreover, κ−1
x must not exceed
−1 2
the standard photon contribution κγ ≈ 1 g/cm , apart from perhaps
a factor of order unity. Taking a typical solar interior temperature
of 1 keV, these requirements exclude the shaded region in Fig. 1.1.
The dashed line marks the parameters where the mfp is of order the
solar radius.
Anomalous
Stellar Energy Losses
Bounded by Observations
23
24 Chapter 2
The escape velocity from our galaxy is about 500 km s−1 . However, in a galactic
4
spiral arm the interstellar medium is dense enough to dissipate a SN explosion which
is only able to blow a hole into the interstellar material.
Anomalous Stellar Energy Losses Bounded by Observations 27
Fig. 2.3. Color magnitude diagram for the globular cluster M3 according
to Buonanno et al. (1986), based on the photometric data of 10,637 stars.
Following Renzini and Fusi Pecci (1988) the following classification has been
adopted for the evolutionary phases. MS (main sequence): core hydrogen
burning. BS (blue stragglers). TO (main-sequence turnoff): central hydro-
gen is exhausted. SGB (subgiant branch): hydrogen burning in a thick shell.
RGB (red-giant branch): hydrogen burning in a thin shell with a growing
core until helium ignites. HB (horizontal branch): helium burning in the
core and hydrogen burning in a shell. AGB (asymptotic giant branch): he-
lium and hydrogen shell burning. P-AGB (post-asymptotic giant branch):
final evolution from the AGB to the white-dwarf stage. (Original of the
figure courtesy of A. Renzini.)
28 Chapter 2
Fig. 2.4. Left panels: A typical solar model (Bahcall 1989) as an ex-
ample for a low-mass star which is halfway through its hydrogen-burning
(main-squence) phase. Right panels: Horizontal-branch star (central helium
burning, shell hydrogen burning) with a metallicity Z = 0.004 after about
2.5×107 yr which is about a quarter of the HB lifetime. (Model from the
calculations of Dearborn et al. 1990.)
30 Chapter 2
Fig. 2.5. Main evolutionary phases of low-mass stars. The envelope and core
dimensions depend on the location on the RGB, HB, or AGB, respectively.
The given radii are only meant to give a crude orientation.
Anomalous Stellar Energy Losses Bounded by Observations 31
Fig. 2.6. Evolutionary track of a 0.8 M! star (Z = 0.004) from zero age
to the asymptotic giant branch. The evolutionary phases are as in Fig. 2.3.
(Calculated with Dearborn’s evolution code.)
32 Chapter 2
The core of a red giant reaches its limiting mass when it has become
so hot and dense that helium ignites. Because the nucleus 8 Be which
consist of two α particles (He nuclei) is not stable, He burning proceeds
directly to carbon, 3α → 12 C (“triple-α reaction”), via an intermediate
8
Be state. Because it is essentially a three-body reaction its rate de-
pends sensitively on ρ and T . In a red-giant core helium ignites when
Mc ≈ 0.5 M! with central conditions of ρ ≈ 106 g cm−3 and T ≈ 108 K
where the triple-α energy generation rate per unit mass, #3α , varies ap-
proximately as ρ2 T 40 . This steep temperature dependence allows one
to speak of a sharp ignition point even though there is some helium
burning at any temperature and density.
The helium core of a red giant is like a powder keg waiting for a
spark. When the critical temperature is reached where #3α exceeds
the neutrino losses a nuclear runaway occurs. Because the pressure
is mainly due to degenerate electrons the energy production at first
does not lead to structural changes. Therefore, the rise in temperature
is unchecked and feeds positively on the energy generation rate. As
this process continues the core expands nearly explosively to a point
where it becomes nondegenerate and the familiar self-regulation by the
gravitational negative specific heat kicks in (Chapter 1). The “explosion
energy” is absorbed by the work necessary to expand the core from
about 106 g cm−3 to about 104 g cm−3 . The core temperature of the final
configuration remains at about 108 K because of the steep temperature
dependence of #3α which allows only for a narrow range of stationary
burning conditions.
The final configuration with a helium-burning core and a hydrogen-
burning shell (Fig. 2.5) is known as a horizontal-branch (HB) star, a
term which is justified by the location of these objects in the color-
magnitude diagram Fig. 2.3. Note that the overall luminosity has de-
creased by the process of helium ignition (Fig. 2.6) because of the core
expansion which lowers the gravitational potential at the core edge and
thus the temperature in the hydrogen-burning shell which continues to
be regulated by the core mass and radius. The total luminosity of an
HB star is given to about 1/3 by He burning and 2/3 by hydrogen shell
burning (Fig. 2.4, right panels).
Because helium ignition is an almost explosive process on dynamical
time scales it is known as the “helium flash.” For the same reason, a
realistic numerical treatment does not seem to exist (for a review, see
Iben and Renzini 1984). There are two main problems. First, normal
34 Chapter 2
remaining for an HB star after mass loss on the RGB will still depend
on its initial value, rendering this a rather natural possibility. How-
ever, it implies that the globular clusters of our galaxy did not form
at the same time—an age spread of several Gyr is required with the
older clusters predominantly at smaller galactocentric distances (Lee,
Demarque, and Zinn 1994 and references therein). Apart from some
anomalous cases such an age spread appears to be enough to explain
the second parameter phenomenon.
Returning to the inner structure of an HB star, it evolves quietly
at an almost fixed total luminosity (Fig. 2.6). The core constitutes
essentially a “helium main-sequence star.” Because its inner core is
convective it dredges helium into the nuclear furnace at the very center,
leaving a sharp composition discontinuity at the edge of the convective
region (Fig. 2.4). After some 12 C has been built up, 16 O also forms. At
the end of the HB phase, the helium core has developed an inner core
consisting of carbon and oxygen.
Fig. 2.7. Ring Nebula in Lyra (M57), a planetary nebula. (Image courtesy
of Palomar/Caltech.)
For stars which start out sufficiently large the mass loss on the
AGB can be so dramatic that one may speak of the ejection of the
entire envelope. It forms a large shell of gas which is illuminated by
its central star, the newborn white dwarf. Such systems are known as
“planetary nebulae” (Fig. 2.7).
Fig. 2.8. The Crab nebula, remnant of the supernova of A.D. 1054. (Image
courtesy of the European Southern Observatory.)
At this point a shock wave forms at the edge of the core and moves
outward. The implosion can be said to be reflected and thus turned into
an explosion. In practice, it is difficult to account for the subsequent
evolution as the shock wave tends to dissipate its energy by dissociating
iron. Currently it is thought that a revival of the stalled shock is needed
and occurs by neutrinos depositing their energy in the “hot bubble”
below the shock. This region has a low density yet high temperature,
and thus a high entropy per baryon by common astrophysical standards.
Within about 0.3 s after collapse the shock has moved outward and
ejects the entire overburden of the mantle and envelope. This course
of events is the scenario of a type II supernova (SN) explosion.
What remains is an expanding nebula such as the Crab (Fig. 2.8)
which is the remnant of the SN of A.D. 1054, and a central neutron
star (radius about 10 km, mass about 1 M! ) which often appears in
the form of a pulsar, a pulsating source of radiation in some or all
electromagnetic wave bands. The pulsed emission is explained by a
complicated interplay between the fast rotation and strong magnetic
fields (up to 1012 −1013 G) of these objects.
Returning to the moment after collapse of the iron core, it is so dense
(nuclear density and above) and hot (temperature of several 10 MeV),
Anomalous Stellar Energy Losses Bounded by Observations 39
that even neutrinos are trapped. Therefore, energy and lepton num-
ber are lost approximately on a neutrino diffusion time scale of several
seconds. The neutrinos from stellar collapse were observed for the first
and only time when the star Sanduleak −69 202 in the Large Magel-
lanic Cloud (a small satellite galaxy of the Milky Way) collapsed. The
subsequent explosion was the legendary SN 1987A.
After the exhaustion of hydrogen, massive stars move almost hori-
zontally across the Hertzsprung-Russell diagram until they reach their
Hayashi line, i.e. until they have become red supergiants. However, sub-
sequently they can loop horizontally back into the blue; the progenitor
of SN 1987A was such a blue supergiant.
The SN rate in a spiral galaxy like our own is thought to be about
one in a few decades, pessimistically one in a century. Because many
of the ones occurring far away in our galaxy will be obscured by the
dust and gas in the disk, one has to be extremely lucky to witness
such an event in one’s lifetime. The visible galactic SNe previous to
1987A were Tycho’s and Kepler’s in close succession about 400 years
ago. Both of them may have been of type I—no pulsar has been found
in their remnants.7 Of course, in the future it may become possible to
detect optically invisible galactic SNe by means of neutrino detectors
like the ones which registered the neutrinos from SN 1987A.
suming that the WD mass function was peaked around 0.7 M! which
is somewhat larger than the canonical value of 0.6 M! . For the present
purpose, however, the smallness of this difference is taken as a confir-
mation of the standard WD cooling theory.
Anomalous Stellar Energy Losses Bounded by Observations 45
(2.10)
12 16
Taking M = 0.6 M! and an equal mixture of C and O one finds
2
log(dN/dMbol ) = 7
Mbol − 6.84 + log(B3 ) , (2.11)
Fig. 2.11. Luminosity function for 0.6 M! WDs for two values of µ12 (Blin-
nikov and Dunina-Barkovskaya 1994) compared with the observations quoted
in the upper part of Tab. 2.1 (Fleming, Liebert, and Green 1986).
Fig. 2.12. Relative number of 0.6 and 0.8 M! WDs in the two hot tempera-
ture bins of Tab. 2.2 as a function of the anomalous neutrino cooling implied
by a magnetic dipole moment µν (Blinnikov and Dunina-Barkovskaya 1994).
The number of WDs in the temperature range 12,000−40,000 K is normal-
ized to unity. The shaded bands correspond to the observations (Tab. 2.2).
50 Chapter 2
They calculated the distribution for the temperature bins of Tab. 2.2
in analogy to the discussion of neutrino dipole moments in the previous
section. Their results are shown in Fig. 2.13 as a function of α26 where
I have corrected from (F ) = 3, which they used in order to compare
with Raffelt (1986b), to the more appropriate value (F ) = 1.
Fig. 2.13. Relative number of 0.6 M! WDs in the hot and intermediate
temperature bins of Tab. 2.2 as a function of the “fine-structure constant”
α26 = α# /10−26 of pseudoscalar bosons (Blinnikov and Dunina-Barkovskaya
1994). I have corrected from (F ) = 3 to the more appropriate value 1. The
number of WDs in the temperature range 12,000−40,000 K is normalized to
unity. The shaded bands correspond to the observations (Tab. 2.2).
a point at which photon emission from the surface becomes the dom-
inant form of cooling. In Fig. 2.14 the central temperature, surface
temperature, neutrino luminosity, and photon luminosity are shown as
functions of age according to a numerical calculation of Nomoto and
Tsuruta (1987).
Fig. 2.14. Cooling of a neutron star with baryon mass 1.4 M! (gravitational
mass 1.3 M! ) according to Nomoto and Tsuruta (1987). The solid line is
for an equation of state of intermediate stiffness (model FP), the dotted
line for a stiff model (PS). All “nonstandard” effects were ignored such as
nucleon superfluidity, a meson condensate, magnetic fields, and so forth.
Temperatures and luminosities are local, ignoring the gravitational redshift.
56 Chapter 2
1. See the review by Tsuruta (1986) for references to the original literature.
2. Becker and Aschenbach (1995).
3. Becker et al. (1992).
4. Ögelman, Finley, and Zimmermann (1993).
5. Thompson et al. (1991). Finley, Ögelman, and Kiziloğlu (1992).
6. Halpern and Ruderman (1993).
7. Brinkmann and Ögelman (1987).
8. Ögelman and Finley (1993). Anderson et al. (1993).
Anomalous Stellar Energy Losses Bounded by Observations 57
ature then stays almost constant for a long time until photon cooling
from the surface begins to dominate. In this scenario the cooling curve
depends sensitively on the “on switch” set by the occurrence of the di-
rect URCA process anywhere in the star, and by the “off switch” from
superfluidity. As the occurrence of these effects depends on fine points
of the equation of state as well as on the density and thus the stellar
mass there may not be a universal cooling curve for all neutron stars.
Another effect which would accelerate cooling is the occurrence of
a meson condensate (Sect. 4.9.1) because of the increased efficiency of
neutrino emission. Again, this effect depends sensitively on the equa-
tion of state and thus on the density and the stellar mass. The most
recent numerical study of neutron-star cooling with a pion condensate
was performed by Umeda, Nomoto, and Tsuruta (1994).
fast. Therefore, the stars along the red-giant branch (RGB), horizon-
tal branch (HB), and asymptotic giant branch (AGB) in a globular
cluster have almost identical initial masses whence for these phases a
single-star track is practically identical with an isochrone. On the other
hand, the TO region requires the construction of detailed theoretical
isochrones to compare theory and observations and thus to determine
the ages of globular clusters.
In order to associate a certain stellar mass with the TO in a cluster
one needs to know the absolute brightness of the stars at the TO, i.e.
one needs to know the precise distance. All else being equal, the inferred
age varies with the TO luminosity as ∂ log(age)/∂ log LTO = −0.85 or
∂ log(age)/∂VTO = 0.34 (Iben and Renzini 1984). Therefore, a 0.1 mag
error in VTO leads to an 8% uncertainty in the inferred cluster age.
Put another way, because L ∝ (distance)2 a 10% uncertainty in cluster
distances leads to an 18% age uncertainty. This is the main problem
with the age determination of globular clusters.
A particularly useful method to measure the distance is to use
RR Lyrae stars as standard candles. As discussed in Sect. 2.1, their
luminosity is determined almost entirely by their core mass (apart from
a dependence on chemical composition), which in turn is fixed by he-
lium ignition on the RGB which, again, depends only on the chemical
composition and not on the red-giant envelope mass. Therefore, the
brightness of the HB is nearly independent of stellar mass. Conse-
TO
quently, the brightness difference ∆VHB between the HB and the TO
is a distance-independent measure of the TO mass and thus of the clus-
ter age. Moreover, because the color of RR Lyrae stars coincides with
that of the TO region it is not necessary to convert from the measured
brightness with a certain filter (e.g. visual brightness V ) to a bolomet-
ric brightness, i.e. there is no need for a bolometric correction (BC).
Also, RR Lyrae stars are bright and easily identified because of their
TO
pulsations. Therefore, ∆VHB is one of the most important observables
in the color-magnitude diagram of globular clusters (Iben and Renzini
1984; Sandage 1986).
TO
As an example the recent ∆VHB determinations of Buonanno, Corsi,
and Fusi Pecci (1989) in 19 globular cluster are shown in Fig. 2.17 as
a function of metallicity; the logarithmic metallicity measure [Fe/H] is
defined in Eq. (2.15). The best linear fit is
TO
∆VHB = (3.54 ± 0.13) − (0.008 ± 0.078) [Fe/H], (2.13)
so that the HB brightness varies with metallicity almost exactly as the
TO brightness. The measured points are in agreement with a Gaussian
Anomalous Stellar Energy Losses Bounded by Observations 63
distribution about the mean. This result illustrates the level of precision
TO
that presently can be achieved at determining ∆VHB .
In principle, the color of the TO also specifies the location on the
MS and thus the TO mass and cluster age. In practice, the color of
a star is theoretically less well determined than its luminosity because
it depends on the treatment of the photosphere and on the surface
area and thus on the radius which is partly fixed by the treatment
of convection. Still, the TO color is a useful measure for the relative
ages between clusters of identical metallicities. Notably, the distance-
independent color difference ∆(B − V ) between the TO and the base
of the RGB (Fig. 2.16) has been used to establish an age difference of
about (3 ± 1) Gyr between the clusters NGC 288 and 362 (VandenBerg,
Bolte, and Stetson 1990; Sarajedini and Demarque 1990).
In order to constrain the operation of a novel energy-loss mechanism
the brightness of the RGB tip is particularly useful because particle
emission (neutrinos, axions) from a red-giant core delays helium igni-
tion. This delay allows the stars to develop a more massive core and
thus to turn brighter before they become HB stars. Again, the inferred
luminosity of the brightest star on the RGB depends on the distance
whence the most useful observable is the distance-independent bright-
ness difference between the RGB tip and the HB. However, because
the color of RR Lyrae stars and red giants is very different one must
64 Chapter 2
tip
convert to an absolute bolometric brightness difference ∆MHB rather
tip
than using the visual brightness difference ∆VHB .
Fig. 2.18. Evolutionary speed on the RGB for a model with M = 0.8 M! ,
metallicity Z = 10−4 , and initial helium abundance Y = 0.240 (Raffelt and
Weiss 1992). The dashed line is the tangent near the bright end.
Fig. 2.19. Luminosity function of the RGB of the globular cluster NGC 2808
which has [Fe/H] = −1.37 (Fusi Pecci et al. 1990).
Anomalous Stellar Energy Losses Bounded by Observations 65
a) Composition Parameters
MRR = 0.66 − 3.5 Y23 + 0.16 Z13 − ∆RR − 7.3 δMc , (2.20)
Comparing this with Eq. (2.20) at Y23 = −0.01 and δMc = 0 one
finds ∆RR ≈ 0. This is not in contradiction with the brightening of
RR Lyrae stars relative to zero age, it only means that there is a slight
offset relative to the analytic representation Eq. (2.20) derived from the
Sweigart and Gross (1976) calculations. ∆RR shall always refer to the
brightness difference of real RR Lyrae stars relative to Eq. (2.20), it
does not refer to an offset relative to real zero-age HB stars.
(at the RR Lyrae strip) and the RGB tip for which one finds with
Eqs. (2.19) and (2.20)
tip
∆MHB ≡ MRR − Mtip
= 4.24 − 4.4 Y23 + 0.35 Z13 − ∆RR + 4.5 δMc (2.22)
log(tHB /yr) = 8.01 + 0.37 Y23 + 0.06 Z13 − 2.9 δMc . (2.23)
log R = 0.105 + 2.29 Y23 + 0.029 Z13 + 0.33 ∆RR − 0.70 δMc .
(2.24)
However, because there are relatively few stars near the tip on the
RGB (on average about 10 mag−1 in the observed clusters), the bright-
est RG is on average about 0.1 mag below the actual RGB tip. Raffelt
(1990b) estimated the richness of the RGB near the tip for each cluster
on the basis of the first few brightest stars provided by the observa-
tions and thus estimated the expected brightness difference between
the brightest RG and the tip. This yields a linear regression
tip
∆MHB = (4.19 ± 0.03) + (0.41 ± 0.06) Z13 , (2.25)
about 0.13 mag brighter than the fit shown in Fig. 2.20. The slope of
Eq. (2.25) agrees very well with the theoretical expectation Eq. (2.22),
provided that ∆RR does not introduce a large modification.
More recent observations are those of Da Costa and Armandroff
(1990) who also found excellent agreement between the theoretical slope
of the brightness of the RGB tip luminosity as a function of metallicity.
Because the coefficient of the metallicity dependence agrees well
with the predicted value one may restrict a further comparison be-
tip
tween theory and observation to ∆MHB at a given metallicity for which
it is best to use the average value [Fe/H] = −1.48 or Z13 = −0.18
of the globular clusters used in Fig. 2.20. Inserting these values into
Eqs. (2.22) and (2.25) and adding the errors quadratically one finds
4.4 Y23 + ∆RR − 4.5 δMc = 0.06 ± 0.03. (2.26)
If ∆RR = 0.2 mag this result implies that the envelope helium abun-
72 Chapter 2
Fig. 2.21. Number ratio of HB/RGB stars for 15 globular clusters according
to Buzzoni et al. (1983).
With ∆RR ≈ 0.2 mag and δMc ≈ 0 this confirms a primordial helium
abundance of around 23%.
This is in good agreement with the standard values Yenv = 0.23, δMc =
0, and ∆RR ≈ 0.1 mag.
Fig. 2.22. Allowed values for the envelope helium abundance of evolved
globular-cluster stars and of an anomalous core-mass excess at helium igni-
tip
tion. The limits were derived from the observed brightness difference ∆MHB
between the HB and the RGB tip, from the “R-method” (number counts
on the HB vs. RGB), and from the brightness determination of nearby field
RR Lyrae stars (MRR ) by statistical parallaxes and the Baade-Wesselink
method as well as the brightness of RR Lyrae stars in the LMC.
76 Chapter 2
if one adopts Yenv = 0.24 ± 0.02, ∆RR = 0.1 ± 0.1, and δMc =
±0.010 M! . Therefore, R cannot be smaller by much more than about
10% of its standard value. Put another way, the helium-burning life-
time of low-mass stars is determined within about 10% from the ratio
of HB/RGB stars in globular clusters.
With lesser statistical significance this result is corroborated by
number counts of clump giants in open clusters which consist of more
recently formed stars in the galactic disk (Population I). Clump giants
in open clusters correspond to HB stars in globular clusters; instead
of forming a horizontal branch they are concentrated in a clump near
the base of the RGB. Cannon (1970) compared the number of clump
giants in the open cluster M67 with the number of stars per luminos-
ity interval near the MS turnoff and found tHe ≈ 1.5×108 yr, with a
large statistical uncertainty, however, because there were only 5 clump
giants. (Open clusters tend to be much less populous than globular
ones.) Tinsley and Gunn (1976) derived tHe = (1.27 ± 0.29) × 108 yr
from low-mass giants of the old galactic disk population. These results
are in full agreement with Eq. (2.23).
Bands of allowed values for Yenv and δMc are shown in Fig. 2.23 which
is analogous to Fig. 2.22.
From this analysis one infers a best-fit value for the envelope he-
lium abundance which is somewhat low. The primordial abundance
probably is not lower than 22%. The only possibility to reduce the
envelope abundance from this level is by gravitational settling which
is counteracted by convective dredge-up, and perhaps by other effects
that might eject helium into the envelope from the core. Therefore,
Fig. 2.23. Allowed values for the envelope helium abundance of evolved
globular-cluster stars and of an anomalous core-mass excess at helium igni-
tion according to the analysis of Catelan, de Freitas Pacheco, and Horvath
(1995). For comparison see Fig. 2.22.
78 Chapter 2
into question in recent years (e.g. Wheeler, Sneden, and Truran 1989).
It is thought that in these systems the “α elements” (mostly oxygen)
are enhanced relative to iron. RG sequences calculated by Raffelt and
Weiss (1992) with somewhat extreme α enhancements indicated that
the red-giant core mass and luminosity at the helium flash are only
moderately changed. Most of the change that did occur was due to
the reduction of Fe at constant Z because of the α enhancement. Put
another way, if Z13 = [Fe/H] + 1.3 is used as the defining equation for
the above “reduced metallicity” the effect of α enhancements appear
to be rather minimal.
Catelan, de Freitas Pacheco, and Horvath (1995) have taken the
point of view that for enhanced α elements one should rescale the
metallicity according to a recipe given by Chieffi, Straniero, and Salaris
(1991). Put another way, the metallicity parameter of Eq. (2.16) that
was used in the previous sections should be redefined as
Z13 ≡ [Fe/H] + 1.3 + log(0.579 f + 0.421). (2.39)
Here, f is the enhancement factor of the abundance of α elements
relative to the solar value. For f = 3 one finds that Z13 must be
offset by +0.33. Catelan, de Freitas Pacheco, and Horvath (1995) went
through their analysis with this modification, causing the allowed bands
of Fig. 2.23 to be slightly shifted relative to each other. However, the
overall change is small, well within the stated upper limit on δMc .
In summary, no effect has been discussed in the literature that would
cause the predicted core mass at helium ignition to deviate from its
standard value beyond the adopted limit of ±0.025 M! . Therefore, any
new energy-loss mechanism that would cause a significantly larger core-
mass excess would have to be compensated by a hitherto unidentified
other novel effect.
from the number of clump giants in open clusters and from the old
galactic disk population.
In Sect. 1.3.1 it was shown that the main impact of a nonstandard
energy-loss rate on a star is an acceleration of the nuclear fuel con-
sumption while the overall stellar structure remains nearly unchanged.
The temperature dependence of the helium-burning (triple-α reaction)
energy generation rate #3α ∝ T 40 is much steeper than the case of hy-
drogen burning discussed in Sect. 1.3.1 and so the adjustment of the
stellar structure is even more negligible. With L3α the standard helium-
burning luminosity of the core of an HB star and Lx the nonstandard
energy-loss rate integrated over the core, tHe will be reduced by an ap-
proximate factor L3α /(Lx + L3α ). Demanding a reduction by less than
10% translates into a requirement Lx < ∼ 0.1 L3α . Because this con-
straint is relatively tight one may compute both Lx and L3α from an
unperturbed model. If the same novel cooling mechanism has delayed
the helium flash and has thus led to an increased core mass only helps
to accelerate the HB evolution. Therefore, it is conservative to ignore
a possible core-mass increase.
The standard value for L3α is around 20 L! ; see Fig. 2.4 for the
properties of a typical HB star. Because the core mass is about 0.5 M!
the core-averaged energy generation rate is (#3α ) ≈ 80 erg g−1 s−1 . Then
a nonstandard energy-loss rate is constrained by
(#x ) < −1 −1
∼ 10 erg g s . (2.40)
Previously, this limit had been stated as 100 erg g−1 s−1 , overly con-
servative because it was not based on the observed HB/RGB number
ratios in globular clusters. However, in practice Eq. (2.40) does not
improve the constraints on a novel energy-loss rate by a factor of 10
because the appropriate average density and temperature are somewhat
below the canonical values of ρ = 104 g cm−3 and T = 108 K.
For a simple estimate the energy-loss rate may be calculated for
average conditions of the core. Typically, #x will depend on some small
power of the density ρ, and a somewhat larger power of the temperature
T . For the HB star model of Fig. 2.4 the core-averaged values (ρn ) and
(T n ) are shown in Fig. 2.24 as a function of n. The dependence on n
is relatively mild so that the final result is not sensitive to fine points
of the averaging procedure.
In order to test the analytic criterion Eq. (2.40) in a concrete exam-
ple consider axion losses by the Primakoff process. The energy-loss rate
will be derived in Sect. 5.2.1. It is found to be proportional to T 7 /ρ
and to a coupling constant g10 = gaγ /10−10 GeV−1 . For a typical HB
Anomalous Stellar Energy Losses Bounded by Observations 81
Fig. 2.24. Average values of ρn and T n for the HB star model of Fig. 2.4
where ρ4 = ρ/104 g cm−3 and T8 = T /108 K.
3 GN M 2
E=− . (2.41)
7 R
The radius of a low-mass white dwarf (nonrelativistic electrons!) may
be expressed as R = R∗ (M! /M)1/3 with R∗ = 8800 km so that
$ %1/3
Ė G N M! M Ṁ
(#grav ) = − = . (2.42)
M R∗ M! M!
From the numerical sequences of Sweigart and Gross (1978) one finds
that near the helium flash M ≈ 0.5 M! and Ṁ ≈ 0.8×10−15 M! s−1
so that (#grav ) ≈ 100 erg g−1 s−1 . Therefore, one must require (#x ) *
100 erg g−1 s−1 in order to prevent the helium flash from being delayed.
In order to sharpen this criterion one may use results from Sweigart
and Gross (1978) and Raffelt and Weiss (1992) who studied numerically
the delay of the helium flash by varying the standard neutrino losses
with a numerical factor Fν where Fν = 1 represents the standard case.
The results are shown in Fig. 2.25. Note that for Fν < 1 the standard
neutrino losses are decreased so that helium ignites earlier, causing
δMc < 0. It is also interesting that for Fν = 0 helium naturally
ignites at the center of the core while for Fν > 1 the ignition point
moves further and further toward the edge (Fig. 2.26). This behavior is
84 Chapter 2
Table 2.6. Increase of the core mass at helium ignition because of the emis-
sion of pseudoscalars (Raffelt and Weiss 1995).
Fig. 2.27. Increase of the core mass of a red giant at helium ignition due
to the emission of pseudoscalars according to Tab. 2.6 (Raffelt and Weiss
1995).
core-mass increases given in Tab. 2.6 and shown in Fig. 2.27. The re-
quirement that the core not exceed its standard value by more than 5%
reproduces the analytic bound. This example nicely corroborates the
surprising precision of the simple criterion Eq. (2.43).
Another important case where a detailed numerical study is avail-
able is the emission of neutrinos by the plasma process γ → νν when
they have nonstandard magnetic dipole moments µν . The emission
rates are derived in Sect. 6.5.5 and the simple criterion Eq. (2.43) is
applied in Sect. 6.5.6. It yields a limit µ12 < −12
∼ 2 where µ12 = µν /10 µB
with the Bohr magneton µB = e/2me . The numerical variation of the
core mass with µν is shown in Fig. 2.28 according to Raffelt and Weiss
(1992). An analytic approximation is
& '
3/2
δMc = 0.025 M! (µ212 + 1)1/2 − 1 − 0.17 µ12 . (2.45)
µν < −12
∼ 3×10 µB , (2.46)
Fig. 2.28. Increase of the core mass of a red giant at helium ignition as a
function of an assumed neutrino dipole moment according to Raffelt and
Weiss (1992) with a total stellar mass 0.8 M! and an initial helium abun-
dance of Y = 0.22 or 0.24 (the core-mass increase is found to be the same).
The triangles refer to the metallicity Z = 10−3 , the squares to 10−4 . The
open circles are the corresponding results of Castellani and Degl’Innocenti
(1992) with the same stellar mass, Y = 0.23, and Z = 2×10−4 . The solid
line is the analytic fit Eq. (2.45).
2.6 Summary
What is the bottom line after studying the impact of an anomalous
energy-loss rate on a variety of stellar-evolution phases? While the Sun
remains an interesting object from a pedagogical point of view, its main
use is that of a distant particle source for terrestrial experimentation,
notably to study neutrino properties (Chapters 10 and 12). Supernovae
and their collapsed cores (newborn neutron stars) remain very impor-
tant; they will be discussed in Chapters 11−13. At the present time
the cooling of old neutron stars appears to be a useful laboratory to
study conventional phenomena (Does the direct URCA process occur?
Are there meson condensates or other exotic phases? What is the role
of magnetic fields?). The emission of weakly interacting particles other
than standard neutrinos appears to play a lesser role in view of other
astrophysical limits on their interaction strength.
Then, apart from the supernova arguments to be discussed later,
the most useful and reliable observables to constrain the operation of a
88 Chapter 2
3.1 Introduction
If weakly interacting particles couple directly to electrons, they can
be produced thermally in stellar plasmas without nuclear processes.
For neutrinos, a direct electron coupling was first contemplated after
the universal V −A theory for their interactions had been proposed in
1958. It was realized immediately that such a coupling would allow
for thermal pair production by bremsstrahlung e− (Z, A) → (Z, A)e− νν
89
90 Chapter 3
Fig. 3.1. Compton processes for photon scattering as well as for axion and
neutrino pair production (“photoneutrino process”). In each case there is
another amplitude with the vertices interchanged.
Zuber 1980)
! "
16 ŝ + 1 2 (ŝ2 − 6 ŝ − 3)
σ = σ0 + + ln(ŝ) . (3.1)
(ŝ − 1)2 ŝ2 (ŝ − 1)3
√
Here, ŝ ≡ s/m2e with s the CM (center of mass) energy.14 This cross
section is shown in Fig. 3.2 as a function of the CM photon energy ω.
For ω & me the CM frame is the electron rest frame, ŝ → 1, and one
recovers the Thomson cross section σ = 83 σ0 . For ω ' me one has
ŝ ' 1 and so σ = (σ0 /ŝ) [2 ln(ŝ) + 1] = (πα2 /ω 2 ) [ln(2ω) + 14 ] because
in this limit s = (2ω)2 .
The standard Compton cross section can also be used to study the
photoproduction of novel low-mass vector particles which couple to
electrons in the same way as photons except that the fine-structure
constant must be replaced by the new coupling α" . Then
σ0 ≡ παα" /m2e with α" ≡ g 2 /4π, (3.2)
a definition that pertains to all bosons which couple to electrons with
a dimensionless Yukawa or gauge coupling g.
If novel vector bosons such as “paraphotons” exist (Holdom 1986)
they likely couple to electrons by virtue of an induced magnetic moment
rather than by a tree-level gauge coupling (Hoffmann 1987),
Lint = (g/4me ) ψ e σµν ψe F µν , (3.3)
where g is a dimensionless effective coupling constant and F the para-
14
The square of the CM energy is s = (P + K)2 with P = (E, p) and K = (ω, k)
the four-vectors of the initial-state electron and photon, respectively. The CM frame
is defined by p = −k so that s = (E + ω)2 = [(m2e + ω 2 )1/2 + ω]2 with ω the initial-
state photon energy in the CM frame. In the frame where the target electron is at
rest (p = 0) one finds s = 2ωme + m2e where now ω is the photon energy in the
electron frame. Thus, with ω the photon energy in the respective frames,
# √
1
ω 2 (ŝ − 1)/ ŝ in the CM frame,
= 1
me (ŝ − 1) in the electron rest frame.
2
Particles Interacting with Electrons and Baryons 93
photon field tensor. In the nonrelativistic limit this yields the same to-
tal cross section as the interaction with pseudoscalars Eq. (3.6) below.
This is seen if one compares the matrix elements between two electron
states i and f for the two cases. For paraphotons (momentum k, polar-
ization vector !) it is g (f |eik·r (k × !) · σ|i* while for pseudoscalars it is
g (f |eik·r k · σ|i*. After an angular average the two expressions are the
same. Of course, one must account for the two paraphoton polarization
states by an extra factor of 2.
Fig. 3.2. Total cross section for the Compton process with a final-state
vector, scalar, or pseudoscalar boson according to Eqs. (3.1), (3.5), and
(3.9), respectively, with σ0 defined in Eq. (3.2) and ω the CM initial photon
energy.
3.2.2 Scalars
Grifols and Massó (1986) studied the stellar emission of scalars φ which
couple according to
Lint = g ψ e ψe φ. (3.4)
Integrating their differential cross section I find
! "
−16 1 − 3ŝ (ŝ + 3)2
σ = σ0 + + ln(ŝ) (3.5)
(ŝ − 1)2 2ŝ2 (ŝ − 1)3
shown in Fig. 3.2 (dashed line). For small and large photon energies
this is half the cross section for massless vector bosons which have
two polarization degrees of freedom. For intermediate energies the two
results are not related by a simple factor.
94 Chapter 3
3.2.3 Pseudoscalars
Next, turn to the photoproduction of low-mass pseudoscalars φ which
couple to electrons by the interaction
Lint = (1/2f ) ψ e γµ γ5 ψe ∂ µ φ or Lint = −ig ψ e γ5 ψe φ, (3.6)
where f is an energy scale and g a dimensionless coupling constant.
Both interaction laws yield the same Compton cross section with the
identification g = me /f (see the discussion in Sect. 14.2.3).
If a pseudoscalar (frequency ω, wavevector k) is emitted in a transi-
tion between the nonrelativistic electron states |i* and |f *, the matrix
element is
g 1 $ %% ik·r % &
%
Mpseudoscalar = √ f %e σ · k %i . (3.7)
2me 2ω
This is to be compared with the corresponding matrix element for pho-
ton transitions,
−e 1 $ %% ik·r % &
Mphoton = √ f %e [2! · p + σ · (k × !)]%%i , (3.8)
2me 2ω
with the photon polarization vector ! and the electron momentum oper-
ator p. Therefore, transitions involving pseudoscalars closely compare
with photonic M1 transitions, a fact that was used to scale nuclear or
atomic photon transition rates to those involving axions (Donelly et al.
1978; Dimopoulos, Starkman, and Lynn 1986a,b).
The relativistic Compton cross section for massive pseudoscalars
was first worked out by Mikaelian (1978). The most general discussion
of the matrix element was provided by Brodsky et al. (1986) and by
Chanda, Nieves, and Pal (1988) who also included an effective photon
mass relevant for a stellar plasma. For the present purpose it is enough
to consider massless photons and pseudoscalars. With σ0 as defined in
Eq. (3.2) one finds
' (
ln(ŝ) 3ŝ − 1
σ = σ0 − , (3.9)
ŝ − 1 2ŝ2
shown in Fig. 3.2 (dotted line).
For ω ' me one finds σ = (παα" /2ω 2 ) [ln(2ω) − 34 ], similar to the
scalar case. For ω & me , however,
σ = 43 σ0 (ω/me )2 , (3.10)
so that the cross section is suppressed at low energies. This reduction
is related to the M1 nature of the transition.
Particles Interacting with Electrons and Baryons 95
GF
Hint = √ ψ e γµ (CV − CA γ5 )ψe ψ ν γ µ (1 − γ5 )ψν , (3.11)
2
where GF is the Fermi constant, and the dimensionless couplings CV
and CA are given in Appendix B. Of course, in the early sixties neutral
currents were not known—one used Fierz-transformed charged currents
which gave CV = CA = 1. For general CV ’s and CA ’s the cross section
was first calculated by Dicus (1972) who found15 (Fig. 3.3)
) *
σ = σ0 (CV2 + CA2 ) σ̂+ − (CV2 − CA2 ) σ̂− ,
σNR = σ0 6
35
(CV2 + 5CA2 ) (ŝ − 1)4
= σ0 96
35
(CV2 + 5CA2 ) (ω/me )4 . (3.13)
Fig. 3.3. Dimensionless total cross sections σ̂+ and σ̂− for the photoneutrino
process γe− → e− νν according to Eq. (3.12) with ω the initial photon energy
in the CM frame.
initial- and final-state electrons have the same momentum, reducing the
calculation to an average of the Pauli blocking factor over all electrons
, -
1 + 2 d3 p 1 1
Fdeg = 1 − (E−µ)/T , (3.16)
ne (2π)3 e(E−µ)/T + 1 e +1
where µ is the electron chemical potential and E 2 = m2e + p2 . Then,
1 +∞ ex
Fdeg = p E dE , (3.17)
n e π 2 me (ex + 1)2
where x ≡ (E − µ)/T . For degenerate conditions the integrand is
strongly peaked near x = 0 so that one may replace p and E with the
values pF and EF at the Fermi surface (x = 0), and one may extend the
lower limit of integration to −∞. The integral then yields T so that
Fdeg = 3EF T /p2F , (3.18)
where ne = p3F /3π 2 was used.
Returning to the nonrelativistic, nondegenerate limit note that the
cross sections are of the form σ = σ∗ (ω/me )p so that
σ∗ ne T p+4 + ∞ xp+3 (p + 3)! ζp+4 σ∗ ne T p+4
Q= dx = . (3.19)
π 2 mpe 0 ex − 1 π2 mpe
Here, ζn = ζ(n) is the Riemann zeta function which shall be set equal
to unity.16 In a medium of mass density ρ the electron density is ne =
Ye ρ/mu where Ye is the electron number fraction per baryon and mu
the atomic mass unit. Therefore, the energy-loss rate per unit mass is
(p + 3)! Ye σ∗ T p+4
-= . (3.20)
π2 mu mpe
The average energy of the photons which are converted into weakly
interacting particles is
(ω* = (p + 3) T. (3.21)
For p = 0 one recovers (ω* = 3T for the average energy of blackbody
photons.17
16
ζn → 1 rather quickly with increasing n; for example ζ4 = π 4 /90 ≈ 1.082.
Therefore, the error is small if one takes ζn = 1 which corresponds to using a
Maxwell-Boltzmann rather than a Bose-Einstein distribution. They differ at small
ω where an effective photon mass in the plasma should be taken into account anyway.
Therefore, at the crude level of accuracy where one uses massless photons nothing
is gained by using the Bose-Einstein distribution.
17
With a Bose-Einstein distribution it is (ω*/T = 3 ζ4 /ζ3 = π 4 /30ζ3 ≈ 2.70.
98 Chapter 3
8αα" Ye T 4
-scalar = = α" 5.7×1029 erg g−1 s−1 Ye T84 , (3.22)
π mu m2e
160 αα" Ye T 6
-pseudo = = α" 3.3×1027 erg g−1 s−1 Ye T86 . (3.23)
π mu m4e
The average energy is (ω* ≈ 5T or (ω*/me ≈ 0.08 T8 .
Finally, turn to neutrino pair production for which the NR cross
section was given in Eq. (3.13). The energy-loss rate is
, -8
96 α G2F m6e T
-νν = (CV2 + 5CA2 ) 4
Ye
π mu me
= (CV2 + 5CA2 ) 0.166 erg g−1 s−1 Ye T88 . (3.24)
Fig. 3.4. Pair annihilation processes for the production of neutrino pairs or
new bosons where a second amplitude with the vertices interchanged is not
shown.
stellar energy-loss rate in detail and found that it dominates the other
neutrino emission processes only in such regions of temperature and
density where the overall neutrino luminosity is very small. Therefore,
free-bound transitions do not seem to be of practical importance as a
stellar energy-loss mechanism.
3.5 Bremsstrahlung
3.5.1 Nondegenerate, Nonrelativistic Medium
The last emission process to be discussed is bremsstrahlung (Fig. 3.5)
where an electron emits a boson or a neutrino pair when scattering
off the Coulomb field of a nucleus. Conceptually, bremsstrahlung is
closely related to the Compton process (Fig. 3.1) because in both cases
the electron interacts with electromagnetic field fluctuations of the am-
bient medium which have nonvanishing power for all wavenumbers and
frequencies. The Compton process corresponds to wavevectors which
satisfy the photon dispersion relation so that a real (on-shell) excita-
tion is absorbed. However, in a typical stellar plasma there is more
power in the “off-shell” electromagnetic field fluctuations associated
with the charged particles.18 Moreover, degeneracy effects do not sup-
press bremsstrahlung at high densities, in contrast with the Compton
process.
which is about 0.12 at the center of the Sun and 0.17 in the cores of
horizontal-branch (HB) stars.
Ignoring screening effects, the energy-loss rate per unit mass is
0 2 √ 3
- = α" 5.9×1022 erg g−1 s−1 T82.5 Ye ρ Yj Zj2 + Zj / 2 , (3.30)
j
8 −3
where T8 = T /10 K, ρ is in g cm , and Yj = Xj /Aj is the number
fraction of nuclear species j relative to baryons while Xj is the mass
fraction, Aj the mass number.
19
Note that e− e− → e− e− γ vanishes to lowest order because two particles of
equal mass moving under the influence of their Coulomb interaction do not produce
a time-varying electric dipole moment because their center of mass and “center of
charge” coincide. However, the emission of pseudoscalars corresponds to M1 rather
than E1 transitions and so it is not suppressed.
Particles Interacting with Electrons and Baryons 103
Grifols, Massó, and Peris (1989) have worked out the bremsstrah-
lung rate for a scalar boson; in this case e− e− collisions can be ignored
relative to electron-nucleus scattering. They found in the nonrelativis-
tic and nondegenerate limit, ignoring screening effects which are small,
2α2 α" ne T 1/2 0 Xj Zj2
-= a
π 3/2 mu m3/2
e j Aj
0 Xj Zj2
= α" 2.8×1026 erg g−1 s−1 T80.5 Ye ρ , (3.31)
j Aj
by ne = p3F /3π 2 . This also implies that |q|2 ≈ |p1 − p2 |2 ≈ 2p2F (1 − c12 )
where c12 is the cosine of the angle between p1 and p2 . With these
approximations and the velocity at the Fermi surface βF ≡ pF /EF =
pF /(m2e + p2F )1/2 one finds
, -
π 2 α2 α" T 4 0 2
Q= n j j F,
Z (3.33)
15 m2e j
where
+ +
dΩ2 dΩa (1 − βF2 ) [2 (1 − c12 ) − (c1a − c2a )2 ]
F =
4π 4π (1 − c1a βF ) (1 − c2a βF ) (1 − c12 )(1 − c12 + κ2 )
(3.34)
with κ2 ≡ kS2 /2p2F .
For a single species of nuclei with charge Ze and atomic weight A
the energy-loss rate per unit mass is
π 2 α2 α" Z 2 T 4
-= F
15 A mu m2e
Z2 4
= α" 1.08×1027 erg g−1 s−1 T F, (3.35)
A 8
where again T8 = T /108 K. Because F is of order unity for all condi-
tions, the bremsstrahlung rate mostly depends on the temperature and
chemical composition, and - is not suppressed at high density.
Expanding Eq. (3.34) in powers of βF for nonrelativistic or partially
relativistic electrons one finds
' ( ! ' ( "
2 2 + κ2 2 + 5κ2 2 + κ2 2 2
F = ln + ln − β + O(βF4 ).
3 κ2 15 κ2 3 F
(3.36)
Therefore, in contrast to the nondegenerate calculation this expression
would diverge in the absence of screening.
Another approximation can be made if one observes that Coulomb
scattering is mostly forward, i.e. the main contribution to the integral
is from c12 ≈ 1 which implies c1a ≈ c2a . With c2a = c1a only in the
denominator one obtains
' ( ! ' ( "
2 2 + κ2 2 + 3κ2 2 + κ2
F = ln + ln − 1 f (βF ) (3.37)
3 κ2 6 κ2
with
' (
3 − 2βF2 3 (1 − βF2 ) 1 + βF
f (βF ) = − ln . (3.38)
βF2 2βF3 1 − βF
The function f (βF ) is 0 at βF = 0 and rises monotonically to 1 for
Particles Interacting with Electrons and Baryons 105
The factors F+ and F− are of order unity. Dicus et al. (1976) de-
rived analytic expressions in terms of a screening scale and the Fermi
velocity of the electrons. In fact, because F− is always much smaller
than F+ and CV2 − CA2 is much smaller than CV2 + CA2 , the “minus”
term may be neglected entirely. Therefore, the Dicus et al. (1976) re-
sult is identical with that of Festa and Ruderman (1969). Either one
is correct only within a factor of order unity because Eq. (6.61) was
used as a screening prescription with the Thomas-Fermi wave number
as a screening scale. However, in a degenerate medium electrons never
dominate screening. The most important effect is from the ion corre-
lations which, in a weakly coupled plasma (Γ < ∼ 1), can be included by
Eq. (6.72) with the Debye scale ki of the ions as a screening scale. While
it is easy to replace kTF with ki in these results, the modification of the
Coulomb propagator according to Eq. (6.72) cannot be implemented
without redoing the entire calculation.
A systematic approach to include ion correlations (i.e. screening
effects) was pioneered by Flowers (1973, 1974) who showed clearly how
to separate the ion correlation effects in the form of a dynamic structure
factor from the matrix element of the electrons and neutrinos. This
approach also allows one to include lattice vibrations when the ions
form a crystal in a strongly coupled plasma. In a series of papers Itoh
and Kohyama (1983), Itoh et al. (1984a,b), and Munakata, Kohyama,
and Itoh (1987) followed this approach and calculated the emission rate
for all conditions and chemical compositions.
As an estimate, good to within a factor of order unity, one may use
F+ = 1 and F− = 0. Moreover, inspired by the axion results one can
guess a simple expression which can be tested against the numerical
rates of Itoh and Kohyama (1983). I find that F− = 0 and
' (
2 + κ2 κ2
F+ ≈ ln + (3.42)
κ2 2 + κ2
that was used in a numerical study of axion emission from red giants
by Raffelt and Weiss (1995); details of how it was constructed can be
found there.
At T = 108 K the compound rate (solid line) coincidentally is almost
independent of density. The requirement - < −1 −1
∼ 10 erg g s then yields
a constraint α" <
∼ 10
−26
for any density whence the helium-burning life-
time of HB stars as well as the core mass at helium ignition yield an
almost identical constraint. In detail one needs to apply the helium-
ignition argument at an average core density of 2×105 g/cm3 (the cen-
tral density is about 106 g/cm3 ) and at the almost constant tempera-
ture of T = 108 K. The degenerate emission rate is then approximately
α" 2×1027 erg g−1 s−1 so that
The same case was treated numerically by Raffelt and Weiss (1995)
who implemented the compound energy-loss rate of Fig. 3.6 with vary-
ing values of α" in several red-giant evolutionary sequences. The re-
sults of this work have been discussed in Sect. 2.5.2 where it was used
as one justification for the simple 10 erg g−1 s−1 energy-loss constraint
that was derived there. This detailed numerical study yielded the same
limit Eq. (3.44) on a pseudoscalar Yukawa coupling to electrons.
1.0×10−26 Galactic age inferred from break Bremsstrahlung (degenerate) Sect. 2.2.4
in white-dwarf luminosity function e + 12 C (16 O) → 12 C (16 O) + e + a
One may also consider the Yukawa coupling of scalar (vector) bosons
to baryons (Grifols and Massó 1986; Grifols, Massó, and Peris 1989b).
In this case one may use the Compton process γ + 4 He → 4 He + φ on
a helium nucleus for which the emission rate is given mutatis mutandis
by the same formula as for nonrelativistic electrons. Assuming the
same coupling to protons and neutrons the emission rate is coherently
enhanced by a factor 42 . Moreover, in Eq. (3.22) one must replace
Ye with YHe (number of 4 He nuclei per baryon) which is 14 for pure
helium, me is to be replaced by mHe ≈ 4mu , and α → 4α to account
for the coherent photon coupling. Pulling these factors together one
finds - = αφN 0.7×1023 erg g−1 s−1 for pure helium. Because the helium-
burning lifetime argument limits this energy-loss rate to 10 erg g−1 s−1
one finds
αφN <
∼ 1.5×10
−22
(3.47)
as a limit on the coupling of a scalar boson to a nucleon N . Again, for
vector bosons this limit is a factor of 2 more restrictive.
eB <∼ 3×10 ,
−11
(3.50)
according to Eqs. (3.43) and (3.47), respectively. Tests of the equiv-
alence principle (i.e. of a composition-dependent fifth force) on solar-
system scales yield β < −9
∼ 10 (Sect. 3.6.3) so that
eB <∼ 1×10 .
−23
(3.51)
Apparently this is the most restrictive limit on eB that is currently
available.
One may be tempted to apply the limits from the equivalence prin-
ciple also to a leptonic charge eL . However, in this case one has to worry
about the fact that even neutrinos would carry leptonic charges. The
universe is probably filled with a background neutrino sea in the same
way as it is filled with a background of microwave photons. This neu-
trino medium would constitute a leptonic plasma which screens sources
of the leptonic force just as an electronic plasma screens electric charges
(Zisman 1971; Goldman, Zisman, and Shaulov 1972; Çiftçi, Sultansoi,
and Türköz 1994; Dolgov and Raffelt 1995).
Debye screening will be studied in Sect. 6.4.1. For the screening
wave number one finds the expression
+ ∞
kS2 = 2 (eL /π)2 dp fp p (v + v −1 ), (3.52)
0
Processes in a Nuclear
Medium
4.1 Introduction
New particles which couple to nucleons are emitted from ordinary stars
by analogous processes to those discussed in Chapter 3 for electrons.
For example, axions can be produced by the Compton process γp → pa;
the previous results can be easily adapted to such reactions. Presently
I will focus on processes involving neutrinos or axions that are specific
to a nuclear medium, i.e. to supernova (SN) cores or neutron stars.
The main focus of the literature which deals with microscopic pro-
cesses in a nuclear medium was inspired by the problem of late-time
neutron-star cooling. The recent progress of x-ray astronomy has led
to reasonably safe ROSAT identifications of thermal surface emission
from a number of old pulsars (Sect. 2.3). Together with the spin-down
age of these objects one can begin to test neutron-star cooling sce-
narios, notably those that involve novel phases of nuclear matter such
as superfluidity, meson condensates, quark matter, etc. (Shapiro and
Teukolsky 1983; Tsuruta 1992).
117
118 Chapter 4
teracting with a hot nuclear medium, even though these issues are of
paramount importance for a proper quantitative understanding of SN
physics where the interaction of neutrinos with the medium dominates
the thermal and dynamical evolution. My discussion can only be a
starting point for future work that may actually yield some answers to
the questions raised.
Raffelt and Seckel (1995) showed that including mπ causes less than
a 30% reduction of typical neutrino or axion rates for T > 20 MeV. Be-
cause this will be a minor error relative to the dominant uncertainties
the term in square brackets is approximated as [3 − (k̂ · l̂)2 ].
The remaining (k̂ · l̂)2 term is inconvenient without yielding any
significant insights. In a degenerate medium it averages to zero in
expressions such as the axion emission rate while in a nondegenerate
medium it can be as large as about 1.31 (Raffelt and Seckel 1995),
leading to an almost 50% reduction of the emissivity. Still, for the
present discussion I will neglect this term and use
!
|M|2 = 16 (4π)3 απ2 αa m−2
N . (4.3)
spins
While this may seem somewhat arbitrary, it must be stressed that us-
ing an OPE potential to model the nucleon interactions in a nuclear
medium is in itself an approximation of uncertain precision. For the
present discussion a factor of order unity will not change any of the
conclusions.
where P1,2 are the four-momenta of the initial-state nucleons, P3,4 are
for the final states, and Ka is for the axion. The factor 14 is a statistics
Processes in a Nuclear Medium 121
nucleon (baryon) density where the factor 2 is for two spin states. Pauli
blocking factors are omitted: (1 − f3,4 ) → 1.
122 Chapter 4
)∞
Table 4.1. sn = 0 xn s(x) e−x dx for nondegenerate (ND) and degenerate
(D) conditions.
n sn (ND) sn (D)
1 8/5 2ζ3 + 6ζ5 /π 2
2 128/35 31π 4 /315
3 256/21 24ζ5 + (180/π 2 ) ζ7
4 4096/77 82π 6 /315
Processes in a Nuclear Medium 123
explicitly
31π 2 pF T 6
QD 2
a = α a απ ,
945 m2N
−1 −1 −2/3
+D 31
a = αa 1.74×10 erg g s ρ15 6
TMeV , (4.10)
QD 31π 4 −5/2
a
= √ ξ ≈ 1.39 ξ −5/2 . (4.11)
QND
a 1536 2
scale for the proton and neutron Fermi momenta which are pn,p F =
(3π 2 nB )1/3 Yn,p
1/3
. According to a result of Brinkmann and Turner (1988)
the effective coupling is then
αa → αn Yn1/3 + αp Yp1/3 + ( 28 α +
3 0
20
α ) (Yn2/3
3 1
+ Yp2/3 )1/2
$ %
1 |Yn2/3 − Yp2/3 |
× √ 2− 2/3 2/3
. (4.14)
2 2 Yn + Yp
The third term is from np collisions; it has the remarkable feature that it
does not vanish as Yp → 0. For equal couplings (Cn = Cp ) the variation
of the emission rate is shown in Fig. 4.4 as a function of Yp = 1 − Yn .
For all proton concentrations the np contribution dominates.
1 ' (
4
d3 pi
S µν ≡ f1 f2 (1 − f3 )(1 − f4 )
nB i=1 2Ei (2π)3
Here, the Pi are the nucleon four momenta, K = −Ka for axion emis-
sion, and K = −(K1 + K2 ) for neutrino pairs. Thus, K is the energy-
momentum transfer from the radiation (axions or neutrino pairs) to
the nucleons. Because S µν knows about the radiation only through
the energy-momentum δ function it is only a function of K = (ω, k),
apart from the temperature and chemical potentials of the medium.
The slightly awkward definition of the sign of K follows the common
definition of the structure function where a positive energy transfer ω
refers to energy given to the medium.
The energy-loss rates by νν or axion emission are then the phase-
space integrals
$ %2 '
CAN GF d3 k 1 d3 k 2
Qνν = √ nB Sµν N µν (ω1 + ω2 ),
2 2ω1 (2π)3 2ω2 (2π)3
$ %2 '
CN d 3 ka
Qa = nB Sµν Kaµ Kaν ωa . (4.19)
2fa 2ωa (2π)3
Here, it was assumed that both axions and neutrinos can escape freely
from the medium so that final-state Pauli blocking or Bose stimulation
factors can be ignored.
In the nonrelativistic limit the nucleon current in Eqs. (4.1) and
(4.15) reduces to χ† τi χ where χ is a nucleon two-spinor and τi (i =
1, 2, 3) are Pauli matrices representing the nucleon spin operator. Put
another way, in the nonrelativistic limit the axial-vector current rep-
resents the nucleon spin density. Therefore, it has only spatial com-
ponents so that S µν → S ij (i, j = 1, 2, 3). In order to construct the
most general tensorial structure for S ij in an isotropic medium only δij
is available. Recall that in the nonrelativistic limit S µν does not know
about the momentum transfer k because of Eq. (4.5). There is then no
Processes in a Nuclear Medium 129
+ND
νν = 2.4×1017 erg g−1 s−1 ρ15 TMeV
5.5
, (4.23)
One can correct for a nonvanishing value of the pion mass by virtue of
Eq. (4.12).
For a mixture of protons and neutrons the same remarks as in
Sect. 4.2.6 apply. Apart from a small correction the neutrino cou-
pling is isovector (CAp ≈ −CAn ) so that α0 ≈ 0 while the other α’s are
approximately equal (Appendix B). For degenerate conditions, with a
small modification the dependence on the proton concentration is the
same as that shown in Fig. 4.4. Again, the absolutely dominating con-
tribution is from np collisions unless protons are so rare that they are
nondegenerate.22
21
Friman and Maxwell’s (1979) total energy-loss rate is 2/3 of the one found here.
Apparently they did not include the crossterm in the squared matrix element, i.e.
the third term in Eq. (4.2).
22
This conclusion, based on the work of Brinkmann and Turner (1988), is in
conflict with the results of Friman and Maxwell (1979). They found that Qνν was
proportional to the proton Fermi momentum which is relatively small in neutron-
star matter. On the other hand, for small proton concentrations Brinkmann and
Turner’s pF in Eq. (4.24) approaches the neutron Fermi momentum. I am in no
position to decide between these conflicting results.
Processes in a Nuclear Medium 131
For the nondegenerate case s(x) was given in Eq. (4.7), for the degen-
erate one in Eq. (4.9). Then one finds numerically κ̂ND = 0.46 and
κ̂D = 1.53. The respective Γσ ’s were given in Eqs. (4.7) and (4.9).
where 1 refers to the ν for which the mfp is being determined while
2 refers to a ν from the thermal environment (occupation number
f2 ). With the detailed-balance relationship Sσ (ω) = Sσ (−ω) eω/T (see
Eq. 4.43) and writing Sσ (|ω|) = (Γσ /ω 2 ) s(ω/T ) as before one finds
'
−1
∞ x22 s(x1 + x2 )
λ = 3CA2 G2F n B Γσ T dx2 f2 , (4.32)
0 2π 2 (x1 + x2 )2
where xi = ωi /T . Then one may use the previously determined Γσ and
s(x) to find the mfp for given conditions (degenerate or nondegenerate).
The ratio between the inverse mfp’s from elastic scattering and pair
absorption is, ignoring the contribution from CV2 ,
λ−1
pair Γσ ' ∞ x22 s(x1 + x2 )
= dx 2 f 2 . (4.33)
λ−1
scat T 0 2π x21 (x1 + x2 )2
An average with regard to a thermal x1 distribution yields about 0.02
for the dimensionless integral where Fermi-Dirac distributions with
chemical potentials µν = 0 were used. The main figure of merit, how-
ever, is Γσ /T , the ratio of a typical spin-fluctuation rate and the ambi-
ent temperature. In the nondegenerate limit one finds with Eq. (4.7)
- .1/2
Γσ nB ρ 30 MeV
γσ ≡ = 4π 1/2 απ2 5/2 ≈ 16 , (4.34)
T mN T 1/2 ρ0 T
Fig. 4.6. In this process, neutrinos can give or take energy even though
recoil effects for heavy nucleons are small in the elastic scattering pro-
cess νN → N ν.
This reaction is identical with pair absorption with the antineutrino
line crossed into the final state. The corresponding mfp is given by the
same expression as for pair absorption, except that the index 2 now
refers to the final-state ν, the initial-state occupation number f2 is to
be replaced with a final-state Pauli blocking factor (1 − f2 ), and the
energy transfer is ω = ω2 − ω1 ,
However, because the energy transfer can be zero, and because Sσ (ω) ∝
ω −2 , this expression diverges. The ω −2 behavior seemed harmless before
because it was moderated by powers of ω from the phase space of axions
or neutrinos.
The occurrence of this divergence could have been predicted without
a calculation by inspecting Fig. 4.6. If one cuts the intermediate-state
nucleon line, this graph falls into two sub-processes (nucleon-nucleon
scattering and nucleon-neutrino scattering) which are each permitted
by energy-momentum conservation, allowing the intermediate nucleon
in the compound process to go “on-shell.” Therefore, the pole of the
propagator which corresponds to real particles causes a divergence of
the cross section. Physically, the divergence reflects a long-range inter-
action which occurs because the intermediate nucleon can travel arbi-
trarily far when it is on its mass shell.
Still, the inelastic scattering process is an inevitable physical possi-
bility. For nonzero energy transfers its differential rate is given by the
unintegrated version of Eq. (4.35). For a vanishing energy transfer it is
elastic and then its rate should be given by Eq. (4.30).
Processes in a Nuclear Medium 135
where rij ≡ ri −rj . When averaged over a thermal ensemble the second
term will disappear if there are no spatial correlations, and nB = NB /V
so that Sρ (k) = 1.
Processes in a Nuclear Medium 139
For the static spin-density structure function the same steps can be
performed with the inclusion of the spin operators 12 σ i for the individual
nucleons. This takes us to the equivalent of Eq. (4.47)
NB
1 !
*Ψ|s(k) · s(−k)|Ψ+ = ek·rij 14 *Ψ|σ i · σ j |Ψ+ =
V i,j=1
NB NB
1 ! 2 1 !
= 1
*Ψ|σ i |Ψ+ + eik·rij 14 *Ψ|σ i · σ j |Ψ+. (4.48)
V i=1 4 V i,j=1
i!=j
1 ' 2d3 p
Sρ,σ (k) = fp (1 − fp+k ), (4.49)
nB (2π)3
' 8 NB
9
+∞ dω 4 !
Sσ (ω, k) = 1 + σ i · σ j cos(k · rij ) . (4.50)
−∞ 2π 3nB i,j=1
i!=j
then yields
' ' +∞
+∞ dω dω ' +∞
ωSρ (ω, k) = ω dt eiωt *ρ(t, k)ρ(0, −k)+
−∞ 2π −∞ 2π −∞
(4.51)
and a similar expression for Sσ . Under the integral, a partial integration
with suitable boundary conditions allows one to absorb the ω factor,
at the expense of ρ(t, k) → ρ̇(t, k). Because Heisenberg’s equation of
motion informs us that iρ̇ = [ρ, H] with H the complete Hamiltonian
of the system one finds (Sigl 1995b)
' +∞ dω 1 35 6 4
ω Sρ (ω, k) = ρ(k), H ρ(−k) ,
−∞ 2π nB
' +∞ dω 4 35 6 4
ω Sσ (ω, k) = σ(k), H · σ(−k) . (4.52)
−∞ 2π 3nB
)
Here it was used, again, that dω eiωt = 2πδ(t) and ρ(k) ≡ ρ(0, k) and
σ(k) ≡ σ(0, k).
In order to evaluate this sum rule more explicitly one must assume
a specific form for the interaction Hamiltonian. In the simplest case of
a medium consisting of only one species of nucleons one may assume
that H consists of the kinetic energy for each nucleon, plus a general
nonrelativistic interaction potential between all nucleon pairs which
depends on the relative distance and the nucleon spins, i.e.
NB NB
! p2i
1 !
H= + V (rij , σ i , σ j ), (4.53)
i=1 2mN 2 i,j=1
i!=j
where again rij ≡ ri −rj . One can then proceed to evaluate the commu-
tators in Eq. (4.52). By virtue of the continuity equation for the particle
number one can then show (Pines and Nozières 1966; Sigl 1995b)
' +∞ dω k2
ω Sρ (ω, k) = ,
−∞ 2π 2mN
' +∞ dω k2
ω Sσ (ω, k) =
−∞ 2π 2mN
8 NB 5
9
4 ! 6
+ σ(k), V (rij , σ i , σ j ) · σ(−k) . (4.54)
3nB i,j=1
i!=j
For the spin-density structure function one can go one step further
by expressing the most general nonrelativistic interaction potential as
(Sigl 1995b)
' +∞ dω k2
ω Sσ (ω, k) =
−∞ 2π 2mN
8 NB 5
9
4 ! 6 5 6
+ 1 − cos(k · rij ) VijS 1
+ 1 + cos(k · rij )
2
VijT . (4.56)
3nB i,j=1
i!=j
It should be stressed that this quantity is not identical with Sρ,σ (ω, 0).
For example, in Eq. (4.47) for k = 0 the interference term does not
average to zero. The structure function becomes NB2 /V and thus co-
herently enhanced because the momentum transfer is so small that a
target consisting of many particles in a volume V cannot be resolved.
The limit k → 0 is understood such that |k|−1 remains much smaller
than the geometrical dimension V 1/3 of the system.
In the long-wavelength limit the normalization Eq. (4.50) and the
f-sum rule Eq. (4.56) yield for the spin-density structure function
' ' ∞ 8 NB
9
+∞ dω dω 4 !
Sσ (ω) = (1 + e−ω/T )Sσ (ω) = 1 + σi · σj ,
−∞ 2π 0 2π 3nB i,j=1
i!=j
' ' ∞ 8 NB
9
+∞ dω dω 4 !
ω Sσ (ω) = ω (1 − e−ω/T )Sσ (ω) = 3 T
V
2 ij
.
−∞ 2π 0 2π 3nB i,j=1
i!=j
(4.58)
These relations will be of great use to develop a general understanding
of the behavior of Sσ (ω) at high densities. The second column of expres-
sions follows from the first by detailed balance. Because Sσ (ω) ≥ 0 it is
evident that all of these expressions are always positive, independently
of details of the medium interactions.
In a noninteracting medium the operators ρ(t, k) and s(t, k) are
constant so that Sρ,σ (ω) = 2πδ(ω), allowing for scattering (zero energy
transfer), but not for the emission of radiation. This behavior is familiar
from a gas of free particles which can serve as targets for collisons, but
which cannot emit radiation because of energy-momentum constraints.
In an interacting medium the density correlator retains this prop-
erty )because in the long-wavelength limit it depends on ρ(t, k→0) =
V −1 d3 r ρ(t, r) which remains constant. Therefore, even in an inter-
acting medium one expects Sρ (ω) = 2πδ(ω), in agreement with the
finding that the neutrino vector current does not contribute to brems-
strahlung in the nonrelativistic limit relative to the axial-vector current
(Friman and Maxwell 1979). ) 7 B
The relevant quantity for the latter is V −1 d3 r s(t, r) = 12 N i=1 σ i
with σ i the individual nucleon spins. If the evolution of different spins
Processes in a Nuclear Medium 143
is uncorrelated one may ignore the cross terms in the correlator. With
the single-nucleon spin operator σ one finds then
' +∞ 3 4
Sσ (ω) = 1
3
dt eiωt σ(t) · σ(0) . (4.59)
−∞
for the differential axion energy-loss rate (radiation power) per nucleon.
Because of collisions with other nucleons, and because of a spin depen-
dent interaction potential caused by pion exchange, the nucleon spins
evolve nontrivially so that the correlator has nonvanishing power at
ω 0= 0, allowing for axion emission.
The same result can be found with the above methods applied to the
spatial part of the vector current. Easier still, it can be obtained directly
from Eq. (4.61). To this end note that photon emission by an electron
involves nonrelativistically (e/me ) p = ev so that we must substitute
σ̇ → v̇ = a. The role of k (axions) is played by the polarization
vector # (photons) so that k2 = ω 2 must be replaced by #2 = 1. With
(CA /2fa ) → e, using α = e2 /4π, and inserting a factor of 2 for two
photon polarization states completes the translation.
In order to understand the radiation spectrum consider a single
“infinitely hard” collision with σ̇(t) = ∆σ δ(t). The radiation power is
$ %2
dIa CN ω2
= 2
|∆σ|2 . (4.63)
dω 2fa 12π
For photons one obtains the familiar flat bremsstrahlung spectrum
dIγ /dω = (2α/3π)|∆v|2 which is hardened, for axions, by the addi-
tional factor ω 2 from their derivative coupling.
In the form Eq. (4.63) the total amount of energy radiated in a sin-
gle collision is infinite. In practice, collisions are not arbitrarily hard,
and the backreaction of the radiation process on the emitter must be
included. This is not rigorously possible in a classical calculation, it re-
quires a quantum-mechanical treatment. Therefore, a classical analysis
is useful only for the soft part of the spectrum where backreactions can
be ignored, i.e. for radiation frequencies far below the kinetic energy
of the emitter. In a thermal environment, a classical treatment then
appears reasonable for ω < ∼ T.
Next, consider a large random sequence of n hard collisions with a
spin trajectory
n
!
σ̇(t) = ∆σ i δ(t − ti ). (4.64)
i=1
where the first term (the “diagonal” part of the double sum) gives the
total radiation power as an incoherent sum of individual collisions. The
Processes in a Nuclear Medium 145
ω2
F (ω) = . (4.67)
ω 2 + Γ2σ /4
*(∆σ)2 +
Γσ = Γcoll . (4.69)
*σ 2 +
Fig. 4.8. Schematic variation of the axion emission rate per nucleon with
Γσ /T , taking Eq. (4.72) for the spin-density structure function. The dashed
line is the naive rate without the inclusion of multiple-scattering effects, i.e.
it is based on Sσ (|ω|) = Γσ /ω 2 .
Processes in a Nuclear Medium 149
and may even decrease at large densities, although such large values for
Γσ may never be reached in a nuclear medium as will become clear be-
low.
The high-density downturn of the axion emission rate can be in-
terpreted in terms of the Landau-Pomeranchuk-Migdal effect (Landau
and Pomeranchuk 1953a,b; Feinberg and Pomeranchuk 1956; Migdal
1956) as pointed out by Raffelt and Seckel (1991). The main idea is
that collisions interrupt the radiation process. The formation of a radi-
ation quantum of frequency ω takes about a time ω −1 according to the
uncertainty principle and so if collisions are more frequent than this
time, the radiation process is suppressed. Classically, this effect was
demonstrated in the language of current correlators in Sect. 4.6.5.
The impact of the high-density behavior of Sσ (ω) on the neutral-
current neutrino opacity is crudely estimated by the inverse mean free
path given in Eq. (4.35), averaged over a thermal energy spectrum of
the initial neutrino. Moreover, all expressions become much simpler if
one replaces the Fermi-Dirac occupation numbers with the Maxwell-
Boltzmann expression e−ωi /T ; the resulting error is small for nondegen-
erate neutrinos. The relevant quantity is then
3 4 ' ∞ ' ∞
−1
λ ∝ dω1 dω2 ω12 ω22 e−ω1 /T Sσ (ω1 − ω2 ). (4.74)
0 0
One integral can be done explicitly, leaving one with an integral over
the energy transfer alone. With Eq. (4.72) for the structure function
one finds
3 4 ' ∞
−1 Γσ (T 2 + T ω/2 + ω 2 /12) e−ω/T
λ ∝ dω . (4.75)
0 ω 2 + Γ2 /4
This expression is constant for Γσ ' T where Γ = Γσ and thus the
structure function is essentially 2πδ(ω). Indeed, the average scattering
cross section (or mean free path) is not expected to depend on the
density.
For dense media (Γσ ) T ), however, the broadening of Sσ (ω) be-
yond a delta function leads to a decreasing average scattering rate. This
means that at a fixed temperature the medium becomes more trans-
parent to neutrinos with increasing density, even without the impact
of degeneracy effects. This behavior is shown in Fig. 4.9 in analogy to
the axion emission rate Fig. 4.8.
A decreasing cross section is intuitively understood if one recalls
that the nucleon spin is typically flipped in a collision with other nucle-
ons because the interaction potential couples to the spin. Neutrino scat-
tering with an energy transfer ω implies that properties of the medium
150 Chapter 4
Fig. 4.9. Schematic variation of the neutrino scattering rate per nucleon
(axial-vector interaction only) with Γσ /T , taking Eq. (4.72) for the spin-
density structure function. The dashed line is the naive rate without the
inclusion of multiple-scattering effects, i.e. it is based on Sσ (ω) = 2πδ(ω).
23
This statement is based on the assumption that there are no spatial correlations
among the nucleons—possibly a poor approximation in a dense medium. Moreover,
collective oscillations may occur so that it can be too simplistic to treat the medium
as consisting of essentially free, individual nucleons (Iwamoto and Pethick 1982).
Calculations of the long-wavelength structure factor by Sawyer (1988, 1989) in a
specific nucleon interaction model showed a substantial suppression of Sρ (0) and
thus, of the neutrino mfp. Other related works are those of Haensel and Jerzak
(1987) and of Horowitz and Wehrberger (1991a,b) who calculated the dynamic
structure functions within certain interaction models. Unfortunately, these works
do not shed much light on the high-density behavior of the quantities which are of
prime interest to this book such as the axion emission rate from a hot SN core.
Processes in a Nuclear Medium 151
(1989) argued on the basis of the nonlinear sigma model that the combi-
nation of parameters απ2 αa /m2N should remain approximately constant.
Mayle et al. (1989) similarly found that this parameter should remain
somewhere in the range 0.3−1.5 of its vacuum value.
The nucleon effective mass m∗N deviates substantially from the vac-
uum value mN = 939 MeV. This shift has an impact on the kinematics
of reactions and the nucleon phase-space distribution. A calculation of
m∗N has to rely on an effective theory which describes the interaction
of nucleons and mesons. A typical result from a self-consistent rela-
tivistic Brueckner calculation including vacuum fluctuations is shown
in Fig. 4.10. For conditions relevant for a SN core, an effective value as
low as m∗N /mN = 0.5 is conceivable.
Fig. 4.11. The URCA processes. For the modified versions, there are obvious
other graphs with the leptons attached to other nucleon lines, and exchange
amplitudes.
457π 2
QURCA = GF cos2 θC (1 + 3CA2 ) m2N pF,e T 6 , (4.76)
10080
where θC ≈ 0.24 is the Cabbibo angle and CA is the charged-current
axial-vector constant which is −1.26 in vacuum while in nuclear matter
it is suppressed somewhat (Appendix B).
154 Chapter 4
11513π 2 2
Qmod. URCA = α G cos2 θC CA2 pF,e T 8 , (4.77)
120960 π F
ignoring factors of order unity to account for a nonzero mπ and nucleon
correlations. As in the νν processes, only the axial-vector coupling
contributes. Some details of the phase-space integration can be found
in Shapiro and Teukolsky (1983).24
The modified URCA reactions (Fig. 4.11) are closely related to νν
bremsstrahlung and the inelastic scattering process νN N → N N ν dis-
cussed earlier. It is interesting that the rate for the e p n → n n νe pro-
cess does not diverge, in contrast with νN N → N N ν where multiple-
scattering effects had to be invoked to obtain a sensible result. The
divergence was due to the intermediate nucleon going on-shell for a
vanishing energy transfer. In e p n → n n νe the electron has the energy
EF,e , the neutrino T , and because EF,e ) T (degenerate electrons!), the
minimum energy transfer to the leptons is EF,e and thus never zero. Of
course, this is the reason why the modified URCA reaction was in-
voked in the first place: if the intermediate nucleon line is cut, the two
sub-processes are suppressed by energy-momentum constraints.
The direct and modified URCA process should be expressed in terms
of a common structure function applicable to charged-current processes.
One may expect that at high density, spin and isospin fluctuations may
suppress these reactions in analogy to the neutral-current processes
discussed in the previous section. However, the URCA processes have
not been discussed in the literature from this particular perspective.
The equivalent of the modified URCA process for quark matter was
calculated by Iwamoto (1980, 1982), more recently by Goyal and Anand
(1990), and numerically by Ghosh, Phatak, and Sahu (1994) who claim
that Iwamoto’s phase-space approximations can lead to substantial er-
rors in the emission rate.
24
Shapiro and Teukolsky’s phase-space volume involves a factor p3F,e because in
their treatment of the nuclear matrix element there is no cancellation of p2F,e with
1/ω 2 from the nucleon propagator which occurs in a bremsstrahlung calculation.
Processes in a Nuclear Medium 155
where Np± is a plane-wave nucleon state with energy Ep and spin ori-
entation ± relative to kπ , and κ0 is a coupling constant.
In the nonrelativistic limit the axion energy-loss rate corresponding
to the decay N; →N ; a is written as
1 2
' '
d3 k d3 p1 ' d3 p2
Qa = ω f1 (1 − f2 )
2ω(2π)3 (2π)3 (2π)3
!
× 2π δ(E1 − E2 − ω) |M|2 , (4.79)
spins
where k is the axion momentum, p1,2 the nucleon momenta, and f1,2
their occupation numbers. The matrix element, averaged over axion
emission angles (the condensate is not isotropic!) is found to be
$ %2
3 ! 4 C0 − g<A C1
|M|2 = 4
3
A2 κ20
spins 2fa
5 6
× (2π)3 δ 3 (∆p + kπ ) + δ 3 (∆p − kπ ) , (4.80)
π 2 κ20 T 4
Q a = αa A2 = αa A2 1.03×1044 erg cm−3 s−1 T94 , (4.82)
45 |kπ |
Instead of an axion one may also emit a neutrino pair in Fig. 4.12.
The corresponding energy-loss rate was worked out by Muto and Tat-
sumi (1988) who found
2π < 2 2 2 2 κ20 T 6
Qνν = C G m A . (4.83)
945 A F N |kπ |
The gluon in Fig. 4.13 couples to the quarks with the strong fine-
structure constant αs and the gluon propagator involves an effective
mass for which Anand, Goyal, and Iha (1990) used m2g = (6αs /π) p2F
where pF is the Fermi momentum of the quark sea. The axion cou-
pling to quarks is the usual derivative form (Cq /2fa ) ψ q γµ γ5 ψq ∂ µ φ with
q = u, d, s. After a cumbersome but straightforward calculation Anand
et al. found for the energy-loss rate in the degenerate limit
=
62π 2 αs2 pF T 6 Cq2 (I1 + 2I2 ) for qq → qqa,
Qa = × (4.84)
945 (2fa )2 (Cq2 + Cq2" ) I1 for qq # → qq # a,
than nuclear matter (Burrows 1979; Anand, Goyal, and Iha 1990). This
observation has little effect on the neutrino cooling rate of a quark
star because the URCA processes, which are based on charged-current
reactions, dominate for both nuclear or quark matter.
For axions, however, it may seem that the emission rate is much
suppressed relative to nuclear matter. This is certainly true if “ax-
ion” stands for any generic pseudoscalar Nambu-Goldstone boson. The
QCD axion, however, which was introduced to solve the CP problem
of strong interactions (Chapter 14) necessarily has a two-gluon cou-
pling which allows for the gluonic Primakoff effect (Fig. 4.14) which is
analogous to the photon Primakoff effect discussed in Sect. 5.2.
for right-handed ones. After this substitution has been performed all
further effects of mν are of higher order so that one may neglect mν
everywhere except in the global “spin-flip factor.”
162 Chapter 4
27
This conclusion is in contrast to a suggestion that the pair-emission rates
could be relatively important in a SN core (Turner 1992). This statement was
not based on a self-consistent treatment of the medium structure functions, but
rather on a perturbative treatment of the processes involved. The conflict between
Turner’s statement and our finding, which essentially was based on phase-space
considerations, highlights the inadequacy of naive perturbative results in a nuclear
medium.
Chapter 5
Two-Photon Coupling of
Low-Mass Bosons
165
166 Chapter 5
the nucleons, for axions they are the quarks, possibly the charged lep-
tons, and perhaps some exotic heavy quark state not contained in the
standard model. For majorons, such couplings exist to neutrinos, and
possibly to other fermions.
An interaction of the form Eq. (5.2) with charged fermions auto-
matically leads to an electromagnetic coupling of the form Eq. (5.1)
because of the triangle amplitude shown in Fig. 5.1. For one fermion
of charge e and mass m an explicit evaluation leads to the relationship
(e.g. Itzykson and Zuber 1983)
α g α
gaγ = = , (5.3)
π m πf
where m was taken to be much larger than the axion and photon en-
ergies. Remarkably, because g = m/f this coupling does not depend
on the fermion mass, but only on the scale f of symmetry breaking.
In general, one must sum over all possible fermions, taking account
of the appropriate charges which are fractional for quarks, and also of
the proper pseudoscalar coupling to the individual fermions which may
vary from m/f by model-dependent factors of order unity.
Fig. 5.1. Triangle loop for the coupling of a pseudoscalar a (axion) to two
photons.
2
Γa→2γ = gaγ m3a /64π. (5.4)
For pions in the sigma model, the only charged fermion is the proton.
Then gπ◦ γ = α/πfπ and Γπ◦ →2γ = α2 m3π /64π 3 fπ2 = 7.6 eV, in close
agreement with the experimental value. (For subtleties of interpretation
of this result in the context of current algebra see the standard field
theory literature, e.g. Itzykson and Zuber 1983).
168 Chapter 5
The main point for the present discussion is that pseudoscalar mass-
less or low-mass bosons are a natural consequence of certain extensions
of the standard model, and that these particles couple to photons ac-
cording to Eq. (5.1) with a strength
α
gaγ = Caγ , (5.5)
πfa
where fa is the energy scale of symmetry breaking and Caγ is a model-
dependent factor of order unity. (For axions, model-dependent details
of the couplings are discussed in Chapter 14.) In the following I will
explore a variety of consequences arising from this interaction.
This cross section exhibits the usual forward divergence from the
long-range Coulomb interaction. For ma '= 0 it is cut off in vacuum by
the minimum necessary momentum transfer qmin = m2a /2ω (ma % ω);
2 1
the total cross section is then σγ→a = Z 2 gaγ [ 2 ln(2ω/ma ) − 14 ].
In a plasma, the long-range Coulomb potential is cut off by screening
effects; according to Sect. 6.4 the differential cross section is modified
with a factor q2 /(kS2 + q2 ). In a nondegenerate medium the screening
scale is given by the Debye-Hückel formula
4πα " # $
kS2 = nB Ye + Zj2 Yj , (5.7)
T j
where the plasma “mass” of the initial-state photon and the axion mass
were neglected relative to the energy ω. In the limit ω % kS this ex-
2
pression expands as Γγ→a = gaγ ω 2 T /16π which is entirely independent
of the density and chemical composition.
For a stellar plasma, however, this approximation is usually not
justified. Ignoring the plasma frequency for the initial-state photons,
the energy-loss rate per unit volume is
) 2
2 d3 kγ Γγ→a ω gaγ T7
Q= = F (κ2 ), (5.9)
(2π)3 eω/T − 1 4π
|M|2 = gaγ
2
(ωL2 − kL2 ) |k̂L · (kT × !T )|2 . (5.12)
Note that |k̂L · (kT × !T )|2 = |(k̂L × kT ) · !T |2 . Averaging over the two
transverse polarization states yields 12 |k̂L × kT |2 .
If one writes ZL = Z!L ωL2 /(ωL2 − kL2 ) as in Sect. 6.3 and performs the
3
d kL integration in Eq. (5.11) one finds
2 )
gaγ ωa ωL
ΓγT →a = 2
dΩa dωa ZT Z!L |k̂L × kT |2
16 (2π) ωT
% (
δ(ωT + ωL − ωa ) δ(ωT − ωL − ωa )
× + . (5.13)
eωL /T − 1 1 − e−ωL /T
2
gaγ T )
ΓγT →a = 2
dΩa ZT Z!L |k̂L × kT |2 . (5.14)
8 (2π)
In this limit a has the same energy as γT while γL has only provided
momentum.
To finish up, note that in the nondegenerate, nonrelativistic limit
ZT = Z!L = 1. Because ωP % T and ωT = O(T ) we have ωP % ωT
so that kT ≈ ωT . Moreover, kL = kT − ka yielding |k̂L × kT |2 =
|ka × kT |2 /|ka − kT |2 = ωT2 (1 − z 2 )/2(1 − z) = 12 ωT2 (1 + z) where z is the
172 Chapter 5
One finds explicitly -Ei Ej .q = q̂i q̂j T /(1 + q2 /kS2 ) in the classi-
cal limit (Sitenko 1967) where kS2 is the Debye-Hückel wave number of
Eq. (5.7). With this result one easily reproduces the Primakoff transi-
tion rate Γγ→a (Raffelt 1988a).
The language of spectral densities for the electromagnetic field fluc-
tuations forms the starting point for a quantum calculation of the axion
emission rate in the framework of thermal field theory. This program
was carried out in a series of papers by Altherr (1990, 1991), Altherr and
Kraemmer (1992), and Altherr, Petitgirard, and del Rı́o Gaztelurrutia
(1994). Naturally, in the classical limit they reproduced the Primakoff
transition rate ΓγT →a of Eq. (5.8).
In the degenerate or relativistic limit their results cannot be rep-
resented in terms of simple analytic formulae. The most important
astrophysical environment to be used for extracting bounds on gaγ are
low-mass stars before and after helium ignition with a core temper-
ature of about 108 K (Sect. 5.2.5). Altherr, Petitgirard, and del Rı́o
Gaztelurrutia (1994) gave numerical results for the energy-loss rate for
this temperature as a function of density shown in Fig. 5.5 (solid line).
The dashed line is the classical limit Eq. (5.9); it agrees well with the
general result in the low-density (nondegenerate) limit. In the degen-
2
La = g10 1.7×10−3 L& , (5.20)
with L& the solar luminosity and g10 ≡ gaγ × 1010 GeV. (Recalling that
gaγ = (α/πfa ) Caγ this corresponds to fa /Caγ = 2.3×107 GeV.) The
differential flux at Earth is well approximated by the formula
which is shown in Fig. 5.6. The average axion energy is -ωa . = 4.2 keV.
2
The total flux at Earth is Fa = g10 3.54×1011 cm−2 s−1 .
The “standard Sun” is about halfway through its main-sequence
evolution. Therefore, the solar axion luminosity must not exceed its
Fig. 5.6. Axion flux at Earth according to Eq. (5.25) from the Primakoff
conversion of photons in the Sun.
176 Chapter 5
photon luminosity; otherwise its nuclear fuel would have been spent
before reaching an age of 4.5×109 yr. This requirement yields a bound
gaγ < −9 −1
∼ 2.4×10 GeV . (5.22)
Fig. 5.7. Results of the galactic axion search experiments of the Rochester-
Brookhaven-Fermilab (RBF) collaboration (Wuensch et al. 1989) and of
the University of Florida (UF) experiment (Hagmann et al. 1990). The
hatched areas are excluded, assuming a local dark-matter axion density of
5×10−25 g cm−3 = 300 MeV cm−3 . The “axion line” is the relationship be-
tween axion mass and coupling strength for ξ = 1 or E/N = 8/3 according
to Eq. (14.24).
Two-Photon Coupling of Low-Mass Bosons 179
By adjusting the gas pressure within the magnetic field volume one
can make the photon and axion degenerate and thus enhance the tran-
sition rate (van Bibber et al. 1989). This applies, in particular, to
solar axions which have keV energies so that the corresponding pho-
ton dispersion relation in low-Z gases is “particle-like” with the plasma
frequency being the effective mass.
If there is a gradient of the gas density, for example near a star, or
if the gas density and magnetic field strength change in time as in the
expanding universe, suitable conditions allow for resonant axion-photon
conversions in the spirit of the neutrino MSW effect (Yoshimura 1988;
Yanagida and Yoshimura 1988).
For the magnetic conversion of pseudoscalars in the galactic mag-
netic field one must worry about density fluctuations of the interstellar
medium which can be of order the medium density itself. In this case
Eq. (5.29) is no longer valid because it was based on the assumption of
spatial homogeneity of all relevant quantities. Carlson and Garretson
(1994) have derived an expression for the conversion rate in a medium
with large random density variations. They found that it can be sig-
nificantly suppressed relative to the naive result.
Instead of using the solar axion flux one can make one’s own by shining
a laser beam through a long transverse magnetic field region where it
develops an axion component. Then the laser beam is blocked while
the weakly interacting axions traverse the obstacle. In a second magnet
they are back-converted into photons so that one “shines light through
walls” (Anselm 1985; Gasperini 1987; van Bibber et al. 1987). Instead
of a freely propagating beam one may use resonant cavities on either
side of the wall which are coupled by the axion field (Hoogeveen and
Ziegenhagen 1991). Another possibility to improve the sensitivity is to
use squeezed light (Hoogeveen 1990).
An actual experiment was performed by Ruoso et al. (1992) who
used two superconducting magnets of length 440 cm each with a field
strength of 3.7 T. The light beam was trapped in a resonant cavity
in the first magnet, allowing for about 200 traversals; the incident
laser power was 1.5 W. At the end of the second magnet photons were
searched for by a photomultiplier. For an axion mass ma < −3
∼ 10 eV an
upper bound gaγ < 0.7×10−6 GeV−1 was found.
Two-Photon Coupling of Low-Mass Bosons 183
2α2 6 2 2 2 2
7
Lγγ = (E − B ) + 7(E · B) (5.30)
45m4e
(Heisenberg and Euler 1936; see also Itzykson and Zuber 1983).
Fig. 5.8. Vacuum birefringence in the presence of external fields. (a) QED
contribution according to the Euler-Heisenberg γγ interaction. (b) γa or
γπ ◦ oscillations in an external E or B field. (c) Photon birefringence in an
external axion field (axionic domain walls, cosmic axion field). (d) Axion-
mediated contribution in a strong E · B field, e.g. near a pulsar.
184 Chapter 5
2α2 B 2 2α2 B 2
n' = 1 + 7 and n⊥ = 1 + 4 , (5.31)
45 m4e 45 m4e
where
It appears as a source for the arion field on the r.h.s. of Eq. (5.24). Tak-
ing account of the relativistic space-time metric outside of the pulsar,
Mohanty and Nayak (1993) found for the resulting arion field
m−1
a would effectively act as a source for the local a field and so it would
be much smaller.
An inhomogeneous pseudoscalar field configuration represents an
optically active medium (Fig. 5.8c) as was noted, for example, in the
context of axionic domain wall configurations (Sikivie 1984). To low-
est order the dispersion relation for left- and right-handed circularly
polarized light is (e.g. Harari and Sikivie 1992)
k = ω ± 12 gaγ k̂ · ∇a, (5.35)
so that the momentum is shifted by a frequency-independent amount.
The corresponding refractive index is n = 1 ± 12 gaγ ω −1 k̂ · ∇a.
Taking account of the relativistic metric in the strong gravitational
field of a pulsar, Mohanty and Nayak (1993) then found for the time de-
lay between circularly polarized waves which propagate approximately
along the polar axis
8 9
2 2 2 4 B04 R11 Ω2 cos4 α
δt = gaγ . (5.36)
5 575 ω 2 (GN M )6
For the pulsar PSR 1937+21 a polarimetric analysis yields a time
delay 0.37 ± 0.67 µs and thus a 1σ upper limit of δt < 1µs (Klein
and Thorsett 1990). For typical pulsar parameters this allowed Mo-
hanty and Nayak (1993) to place a limit on the arion-photon coupling
of gaγ < −11
∼ 2×10 GeV .
−1
than about 10−9 of a given neutrino species may show up in the form
of decay photons (Fig. 12.9) if the spectral distribution is taken to be
4 <
characterized by T ≈ 30 MeV. Therefore, one finds a limit of g10 ∼
10 or gaγ <
−4 −11 −1
∼ 10 GeV , applicable if the particle mass is below
about 10−10 eV. This is more restrictive than Carlson’s original limit,
and of the same order as the PSR 1937+21 birefringence limit quoted
after Eq. (5.36).
Fig. 5.9. Bounds on the photon coupling gaγ as a function of ma for arbitrary
pseudoscalars; see the text for details. (Adapted from Cameron et al. 1993.)
Chapter 6
6.1 Introduction
Particles are the quantized excitations of certain fields—photons of the
electromagnetic field, electrons of the electron field, and so forth. It
is usually convenient to expand these fields in plane waves character-
ized by frequencies ω and wave vectors k; the excitations of these modes
then exhibit a temporal and spatial behavior proportional to e−i(ωt−k·x) .
The frequency for a given wave number is determined by the disper-
sion relation. Because (ω, k) is a four-vector, and because of Lorentz
invariance, in vacuum the quantity ω 2 − k2 = m2 is the same for all
193
194 Chapter 6
and density, most of the results relevant for particle physics in stars
predate the development of this formalism; they were based on the
old-fashioned tools of kinetic theory. Indeed, for simple issues of dis-
persion or collective effects a kinetic approach seems often physically
more transparent while yielding identical results. At any rate, the fol-
lowing discussion is based entirely on kinetic theory.
only refers to scattering in the forward direction, but that all properties
of the wave and the scatterer are left unchanged. If the medium parti-
cles have a distribution of momenta, spins, etc. the forward scattering
amplitude must be averaged over those quantities, and different species
of medium particles must be summed over.
For a practical calculation it helps to recall that dσ/dΩ = |f (θ)|2
so that |f0 | is the square root of the forward differential cross section.
For example, the Thomson cross section for photons interacting with
nonrelativistic electrons is dσ/dΩ = (α/me )2 |, · ,$ |2 with the polariza-
tion vectors , and ,$ of the initial- and final-state photon. Forward
scattering implies |, · ,$ |2 = 1 so that |f0 | = α/me . The dispersion rela-
tion is then ω 2 = k 2 + ωP2 with the plasma frequency ωP2 = 4πα ne /me .
Of course, the absolute sign of f0 has to be derived from some other
information—for photon dispersion see Sect. 6.3.
The forward scattering amplitude and the refractive index are gen-
erally complex numbers. Physically it is evident that in a medium the
intensity of a beam is depleted as e−z/& . The mean free path is given by
-−1 = σnv where σ is the total scattering cross section, n is the number
density of scatterers, and v is the velocity of propagation. Thus the
amplitude of a plane wave varies as eikz−z/2& . Moreover, the derivation
of the refractive index indicates that the amplitude varies according to
einrefr ωz , yielding k = Re nrefr ω and (2-)−1 = Im nrefr ω. For relativistic
propagation (v = 1) the last equation implies σ(ω) = (4π/ω) Im f0 (ω),
a relationship known as the optical theorem.
For the applications discussed in this book specific interaction mod-
els between the propagating particles and the medium will be assumed
so that it is usually straightforward to calculate the dispersion relation
according to Eq. (6.5). One should keep in mind, however, that nrefr as
a function of ω has a number of general properties, independently of the
interaction model. For example, its real and imaginary part are con-
nected by the Kramers-Kronig relations (Sakurai 1967; Jackson 1975).
where the dispersion relation Eq. (6.9) was used. To first order in ω−ωk
one may use 2ω = 2ωk = ω + ωk which allows one to write
where
(
−1 2ωk − Π$k (ωk ) ∂Π(ω, k) ((
Z ≡ =1− ( . (6.13)
2ωk ∂ω 2 (ω2 −k2 =Π(ω,k)
202 Chapter 6
∇ · E = ρ, ∇ × B − Ė = J,
∇ · B = 0, ∇ × E + Ḃ = 0. (6.15)
∂ · J = ρ̇ − ∇ · J = 0, (6.16)
F µν = ∂ µ Aν − ∂ ν Aµ , (6.18)
A − ∂ (∂ · A) = J, (6.19)
where = ∂ · ∂ = ∂µ ∂ µ = ∂t2 − ∇2 .
31
They are sometimes called “longitudinal plasmons” in contrast to “transverse
plasmons.” In this nomenclature the term “plasmon” refers to any excitation of the
electromagnetic field in a medium while “photon” refers to an excitation in vacuum.
204 Chapter 6
∂ · A = 0, Lorentz gauge,
∇ · A = 0, Coulomb gauge. (6.20)
Φ = ρ, A = J, Lorentz gauge,
−∇2 Φ = ρ, A = JT , Coulomb gauge, (6.21)
particles which constitute the currents move themselves under the in-
fluence of electromagnetic fields. Therefore, the interaction between
fields and currents must be calculated self-consistently. If the fields are
sufficiently weak one may assume that the reaction of the currents to
the fields can be described as a linear response. (For a general review
of linear-response theory in electromagnetism see Kirzhnits 1987.)
In general this statement cannot be made locally in the sense that
the currents at space-time point (t, x) were only linear functions of
A(t, x). Within the restrictions imposed by causality the relation-
ship between fields and currents is nonlocal; for example, a solution
of Maxwell’s equations with prescribed currents requires integrations
over the sources in space and time. After a Fourier transformation,
however, the assumption of a linear response can be stated as
µ
Jind = −Πµν Aν . (6.22)
The polarization tensor Π(K) with K = (ω, k) is a function of the
medium properties.
Besides the induced current there may be an externally prescribed
one Jext which is unrelated to the response of the microscopic medium
constituents to the fields; the total current is J = Jind + Jext . Maxwell’s
equations (6.19) are then in Fourier space
µ
(−K 2 g µν + K µ K ν + Πµν )Aν = Jext . (6.23)
Invariance under a gauge transformation Aν → Aν + Kν α requires that
Πµν Kν = 0. Because the external and total currents are conserved
the induced current is conserved as well, leading to K · Jind = 0 or
Kµ Πµν = 0. Altogether
Kµ Πµν = Πµν Kν = 0 (6.24)
which is an important general property of the polarization tensor.
Considering the Maxwell equations in Coulomb gauge in the ab-
sence of external currents, the transversality of A still implies that it
provides only two wave polarization states, albeit with modified disper-
sion relations due to the presence of Π. With regard to the Φ equation
note that in an isotropic medium the induced charge density ρind must
be a spatial scalar and so can depend only on Φ and the combination
k · A = 0 which is the only available scalar linear in A. Therefore, the
homogeneous equation for Φ is
(k2 + Π00 )Φ = 0. (6.25)
00
Because Π is a function of ω and k this is a wave equation with the
dispersion relation k2 + Π00 (ω, k) = 0. The electric field associated
206 Chapter 6
e± ≡ (0, e± ) (6.28)
−ω 2 + k 2 + πa (ω, k) = 0. (6.32)
It yields the frequency ωk for modes with a given polarization and wave
number. The so-called effective mass is then m2eff = πa (ωk , k). This
expression is different for different polarizations and wave numbers,
and may even be negative.
Generally, an isotropic medium is characterized by three different
response functions because the left- and right-handed circular polariza-
tion states may experience different indices of refraction (Nieves and Pal
1989a,b). Such optically active media are not symmetric under a parity
transformation. For example, a sugar solution changes under a spatial
reflection because the sugar molecules have a definite handedness.
If the medium and all relevant interactions are even under parity the
circular polarization states have the same refractive index. Then one
needs to distinguish only between transverse and longitudinal modes;
one defines πT ≡ π+ = π− and PT = P+ + P− which projects on the
plane transverse to K and U in Minkowski space.
In macroscopic electrodynamics the medium effects are frequently
stated in the form of response functions to applied electric and mag-
netic fields instead of a response to A. The displacement induced by an
applied electric field is D = , E with , the dielectric permittivity. Simi-
larly, the magnetic field is H = µ−1 B for an applied magnetic induction
where µ is the magnetic permeability. For time-varying and/or inho-
mogeneous fields these relationships are understood in Fourier space
where the response functions depend on ω and k.
The magnetic field H and the transverse part of D, characterized
by k · DT = 0, do not have independent meaning (Kirzhnits 1987).
Therefore, among other possibilities one may choose H = B, DT =
,T ET , and DL = ,L EL . In this case ,L ≡ , is the longitudinal and
,T ≡ ,L + (1 − µ−1 ) k 2 /ω 2 the transverse dielectric permittivity.
Particle Dispersion and Decays in Media 209
Fig. 6.3. Contours for v∗ and γ = ωP /T as defined in Eqs. (6.39) and (6.40)
where Ye is the number of electrons per baryon.
Next, with Eq. (6.38) one must solve the transcendental equations
πT,L (ω, k) = ω 2 − k 2 which are explicitly
& '
ω2 − k2 = ωP2 1 + 12 G(v∗2 k 2 /ω 2 ) Transverse,
& '
ω 2 − v∗2 k 2 = ωP2 1 − G(v∗2 k 2 /ω 2 ) Longitudinal. (6.44)
Fig. 6.5. Dispersion relation for transverse modes according to Eq. (6.44).
Fig. 6.6. Dispersion relation for longitudinal modes according to Eq. (6.44).
are
1 + 3T /me Classical,
3 - . 4
k12
3 1 1 + vF
= log −1 Degenerate, (6.49)
ωP2
vF2 2vF 1 − vF
∞ Relativistic.
Then, for k > k1 the four-momentum is space-like, ω 2 − k 2 < 0.
As discussed in Sect. 6.2.2 there is nothing wrong with a space-like
four-momentum of an excitation. In media with electron resonances
such as water or air even (transverse) photons exhibit this behavior
which allows kinematically for their Cherenkov emission e → eγ or
absorption γe → e. In a plasma, transverse excitations are always
time-like and thus cannot be Cherenkov absorbed. Their lowest-order
damping mechanism is Thomson scattering γe → eγ which is not in-
cluded because it is an O(α2 ) effect. Longitudinal excitations with
k > k1 , in contrast, can and will be Cherenkov absorbed by the ambi-
ent electrons, leading to an O(α) damping rate. It corresponds to an
imaginary part of the dispersion relation (an imaginary part of πL ).
In the expression Eq. (6.36) this damping effect corresponds to a
vanishing denominator, essentially to P · K = 0, which occurs when the
intermediate electron in Compton scattering “goes on-shell.” Evidently,
P · K = Eω − p · k can never vanish for k < ω while for k > ω there are
always some electrons, even in a nonrelativistic plasma, which satisfy
this condition. When the phase velocity ω/k becomes of the order of
a typical thermal velocity the number of electrons which match the
Cherenkov condition becomes large, and then the damping of plasmons
becomes strong. Because v∗ measures a typical electron velocity this
occurs for k > ∼ ω/v∗ (Fig. 6.4). Therefore, while nothing dramatic
happens where the dispersion relation crosses the light cone, it fizzles
out near the “electron cone.” For k > ∼ ω/v∗ there are no organized
oscillations of the electrons—longitudinal modes no longer exist.
This damping mechanism of plasma waves was first discussed by
Landau (1946) and is named after him. A calculation in terms of Che-
renkov absorption was performed by Tsytovich (1961). In the classical
limit the Landau damping rate (the imaginary part of the frequency) is
5 ! "3 - .3/2 - .3 ω2
k2 √
ΓL π kD 5 ωP − 52 v∗2 kP2
e− 2k2 = π
D
= e , (6.50)
ωP 8 k 2 v∗ k
where kD = 4πα ne /T is the Debye screening scale. (Note that ωP2 /kD
2
=
T /me = v∗2 /5.) For a given wave number a plasmon must be viewed
Particle Dispersion and Decays in Media 217
2ω 2 (ω 2 − v∗2 k 2 )
ZT = ,
ω 2 [3ωP2 − 2 (ω 2 − k 2 )] + (ω 2 + k 2 )(ω 2 − v∗2 k 2 )
2 (ω 2 − v∗2 k 2 ) ω2
ZL = . (6.52)
3ωP2 − (ω 2 − v∗2 k 2 ) ω 2 − k 2
In each case ω and k are “on shell,” i.e. they are related by the disper-
sion relation relevant for the T and L case, respectively.
Inspection of Eq. (6.52) reveals that ZT is always very close to unity,
as expected for excitations with only a small deviation from a massive-
particle dispersion relation. The contours in Fig. 6.7 confirm that ZT
never deviates from unity by more than a few percent.
218 Chapter 6
Fig. 6.7. Contours for the vertex renormalization factor ZT for transverse
electromagnetic excitations in a medium according to Eq. (6.52).
Fig. 6.8. Modified vertex renormalization factor Z)L for longitudinal electro-
magnetic excitations in a medium according to Eq. (6.53).
(∂t J = 0) are not screened. The magnetic field associated with a sta-
tionary current is the same at a distance whether or not the plasma
is present.
Not so for the electric field associated with a charge. In the static
limit one finds
# ∞
4α
πL (0, k) = dp fp p (v + v −1 ). (6.55)
π 0
4πα ne me 2
kS2 = kD
2
= = ω . (6.58)
T T P
2
At this point one recognizes that kD is independent of the electron
mass, in contrast with the plasma frequency ωP2 . Therefore, it is no
longer justified to ignore the ions or nuclei; they contribute little to
dispersion because of their reduced Thomson scattering amplitude, but
they contribute equally to screening. Therefore, one finds kS2 = kD
2
+ ki2
with
4πα *
ki2 = nj Zj2 , (6.59)
T j
4α 3ω 2
kS2 = kTF
2
= EF pF = 2P . (6.60)
π vF
2 2
However, kD always exceeds kTF so that in a medium of degenerate
electrons and nondegenerate ions the main screening effect is from the
latter. Recall that the Fermi momentum is related to the electron
density by ne = p3F /3π 2 and the Fermi energy is EF = (p2F + m2e )1/2 .
To compare the Thomas-Fermi with the Debye scale take the non-
relativistic limit (kTF /kD )2 = 32 T /(EF − me ). This is much less than 1
or the medium would not be degenerate whence kTF & kD . There-
fore, if the electrons are degenerate and the ions nondegenerate, a test
charge is mostly screened by the polarization of the ion “fluid” because
the electrons form a “stiff” background. Unfortunately, one often finds
calculations in the literature which include screening by the electrons
(screening scale kTF ) but ignore the ions. The resulting error need not
be large because the screening scale typically appears logarithmically
in the final answer (see below).
Screening effects in Coulomb processes are often found to be imple-
mented by a modified Coulomb propagator
1 1
→ , (6.61)
|q|4 (q2 + kS2 )2
kS2 e−kS r
ρ(r) = δ 3 (r) − . (6.64)
4π r
The volume integral of ρ(r) vanishes, giving zero total charge, i.e. com-
plete screening at infinity. If one imagines that only one species of
charged particles is mobile on a uniform background of the opposite
charge, then Eq. (6.64) implies correlations between the mobile species
of n h(r) = −(kS2 /4πr) e−kS r . As expected, Debye screening corresponds
Particle Dispersion and Decays in Media 223
which implies
1 1
→ 2 2 (6.72)
|q| 4 q (q + kS2 )
# +1 (1 − x) f (x)
dx , (6.73)
−1 (1 − x)2
where x is the cosine of the scattering angle of the probe. Here, f (x) is
a slowly varying function which embodies the details of the scattering
or bremsstrahlung process. If this function is taken to be a constant,
the two screening prescriptions amount to the two integrals
# ! "
+1 1 2 + κ2
dx = log ,
−1 (1 − x + κ2 ) κ2
# ! "
+1 (1 − x) 2 + κ2 2
dx = log − , (6.74)
−1 (1 − x + κ2 )2 κ2 2 + κ2
where κ2 ≡ kS2 /2p2 is the screening scale expressed in units of the initial-
state momentum of the probe. Usually, it far exceeds the screening scale
whence κ2 & 1. Then Eq. (6.72) yields a cross section proportional to
log(4p2 /kS2 ) while Eq. (6.61) gives [log(4p2 /kS2 ) − 1].
Thus, if one is only interested in a rough estimate, either screening
prescription and any reasonable screening scale yield about the same
result. For an accurate calculation, however, one needs to identify
the dominant source of screening (for example, the nondegenerate ions
in a degenerate plasma and not the electrons), and the appropriate
moderation of the Coulomb propagator, usually Eq. (6.72).
Particle Dispersion and Decays in Media 227
where explicitly
Mαβ = 4e2ν Z (gαβ + 2,∗α ,β ). (6.77)
Here, Z is the renormalization constant (Sect. 6.2.3 and 6.3.6), P and
P are the ν and ν four-momenta, and , is the plasmon polarization
vector for which one uses the basis vectors of Eqs. (6.27) and (6.28).
34
I closely follow Haft (1993).
35
In this section the term “plasmon” refers to both transverse and longitudinal
electromagnetic excitations in a medium.
228 Chapter 6
Γ = αν Z (ω 2 − k 2 )/3ω, (6.80)
The squared matrix element is of the form Eq. (6.76); for a magnetic
dipole coupling µ of a single flavor one finds
GF
Lint = − √ ψ e γα (CV − CA γ5 )ψe ψ ν γ α (1 − γ5 )ψν . (6.85)
2
The vector-current has the same structure that pertains to the elec-
tron interaction with photons, Lint = −ie ψ e γα ψe Aα . Therefore, af-
ter performing a thermal average over the electron forward scattering
amplitudes the plasmon decay is represented by the Feynman graph
Fig. 6.12 which is identical with Fig. 6.1 with one photon line replaced
by a neutrino pair. As far as the electrons are concerned, photon for-
ward scattering γ → γ is the same as the conversion γ → νν.
present discussion I will not worry any further about the axial-vector
contribution.
The matrix element for the interaction between neutrinos and pho-
tons can then be read from the effective vertex
C V GF
i √ Aα Παβ ψ ν γ α (1 − γ5 )ψν . (6.86)
e 2
The electric charge in the denominator removes one such factor con-
tained in Π which was calculated for photon forward scattering. In
the matrix element, Π is to be taken at the four-momentum K of the
photon. Besides plasmon decay γ → νν, this interaction also allows for
processes such as Cherenkov absorption γν → ν or emission ν → γν.
For the decay of a plasmon with a polarization vector , and four-
momentum K the squared matrix element has the form Eq. (6.76) with
G2F 2
Mαβ = 8 π (gαβ + 2,∗α ,β ). (6.87)
2e2 T,L
Because the plasmon is a propagating mode it obeys its dispersion
relation, i.e. πT,L = ω 2 − k 2 . The decay rate is then
CV2 G2F (ω 2 − k 2 )3
Γ= Z T,L . (6.88)
48π 2 α ω
This result was first derived by Adams, Ruderman, and Woo (1963),
the correct Z for the longitudinal case was first derived by Zaidi (1965).
This equation is understood “on shell” where ω depends on k through
the dispersion relation ω 2 − k 2 = πT,L (ω, k).
The decay rates of Eqs. (6.80), (6.83) and (6.88) of a plasmon with
three-momentum k are then expressed as
ωP2 π̂k
αν Millicharge,
4π
! "2
µ2 ωP2 π̂k
4π Zk
Γk = × Dipole Moment, (6.90)
3 ωk
2 4π
! "3
CV2 G2F ωP2 π̂k
Standard Model,
α 4π
where for Zk and π̂k the T or L value appropriate for the chosen polar-
ization must be used.
1 # k1 n
2 ZL π̂L 1 # k1 2 )
2 ω ZL π̂L
n−1
QL,n = dk k = dk k ,
4ζ3 T 3 0 eω/T − 1 4ζ3 T 3 0 ωP2 eω/T − 1
1 #∞ n
2 ZT π̂T
QT,n = dk k . (6.93)
2ζ3 T 3 0 eω/T − 1
The second equation for QL,n relies on the definition Eq. (6.53), i.e.
ZL = Z)L ω 2 /(ω 2 − k 2 ), and ω 2 − k 2 = πL was used.
The normalization factors were chosen such that QT,n = 1 if the
plasmons are treated as effectively massless particles for the phase-space
integration. Then ZT = π̂T = 1 which is a reasonable approximation
in a nondegenerate, nonrelativistic plasma. In that limit to lowest
order k1 = ωP , Z)L = 1, and πL = ωP2 − k 2 . Therefore, in this limit
QL,n & QT,n . In fact, the longitudinal emission rate is of comparable
importance to the transverse one only in a narrow range of parameters
of astrophysical interest (Haft, Raffelt, and Weiss 1994).
These simple approximations, however, are not adequate for most of
the conditions where the plasma process is important. In Appendix C
the numerical neutrino emission rates are discussed; a comparison be-
tween Fig. C.1 and Fig. 6.3 reveals that the plasma process is important
for 0.3 < <
∼ ωP /T ∼ 30, i.e. transverse plasmons can be anything from
relativistic to entirely nonrelativistic. For a practical stellar evolution
calculation one may use the analytic approximation formula for the
plasma process discussed in Appendix C, based on the representation
of the dispersion relations of Sect. 6.3.
The main issue at stake in this book, however, is nonstandard neu-
trino emission from the direct electromagnetic couplings discussed in
Sect. 6.5. Instead of constructing new numerical emission rate formulae
one uses the existing ones for the standard-model (SM) couplings and
scales them to the novel cases. Numerically, one finds
! "4
Qcharge αν α (4π)2 Q1 10 keV Q1
= 2 2 4 = 0.664 e214 ,
QSM CV GF ωP Q3 ωP Q3
! "2
Qdipole µ2 α 2π Q2 10 keV Q2
= 2 2 2 = 0.318 µ212 , (6.94)
QSM CV GF ωP Q3 ωP Q3
where e14 = eν /10−14 e and µ12 = µ/10−12 µB with µB = e/2me .
Contours for Q1 /Q3 and Q2 /Q3 are shown in Fig. 6.13 according to
Haft, Raffelt, and Weiss (1994). Replacing these ratios by unity in a
practical stellar evolution calculation introduces only a small error.
234 Chapter 6
Fig. 6.13. Contours of Q1 /Q3 and Q2 /Q3 defined by Eq. (6.93) in the plane
defined by the plasma frequency ωP and a “typical” electron velocity v∗
discussed in Sect. 6.3. See Fig. 6.3 for contours of v∗ and ωP in the T -ρ-
plane. (Adapted from Haft, Raffelt, and Weiss 1994.)
that one may employ the simple analytic form Eq. (6.92) of the emission
rate with Qn = 1. The core of HB stars consists at first of helium, later
also of carbon and oxygen, for all of which Ye = 0.5. Then,
5.0 e214 Millicharge,
,x = 1 erg g−1 s−1 × T83 × 0.098 µ212 ρ4 Dipole Moment, (6.95)
0.0127 ρ24 Standard Model,
eν < −14
∼ 2×10 e and µν < −12
∼ 14×10 µB . (6.96)
Of course, for such large dipole moments the core would grow far be-
yond its standard value before helium ignites, causing an additional
acceleration of the HB lifetime. In fact, this indirect impact on the HB
lifetime would be the dominant effect as shown, for example, by the
numerical calculations of Raffelt, Dearborn, and Silk (1989).
From Fig. 6.14 it is clear that the dipole-induced emission rate is
larger for the conditions of the second criterion, based on the helium-
ignition argument where 0ρ1 ≈ 2×105 g cm−3 . According to Eq. (D.12)
the relevant plasma frequency is ωP = 8.6 keV so that Qcharge /QSM ≈
1.2 e214 and Qdipole /QSM ≈ 0.4 µ212 in Eq. (6.94). The average total emis-
sion rate is then given by the standard rate times Fν = 1 + Qj /QSM
where j stands for “charge” or “dipole.” In order to prevent the core
mass at helium ignition from exceeding its standard value by more
than 5% one must require Fν < 3. Then one finds eν < ∼ 1.3×10 e
−14
eν < −14
∼ 2×10 e and µν < −12
∼ 3×10 µB (6.97)
where fe± (p) is the electron and positron phase-space distribution with
P = (E, p) the electron or positron four-momentum.
Instead of a neutrino pair, another photon can be thought of as
being coupled to the electron line in Fig. 6.11 or 6.12, a process which
represents photon forward scattering. Therefore, apart from overall
coupling constants Λαβ V is identical with the electronic contribution to
the photon polarization tensor Παβ studied earlier in this chapter
Λαβ αβ
V = (CV /e) Π . (6.100)
Λαβ
A = 2ieCA ,
αβµ0
Kµ a(ω, k) (6.101)
Λαβ 00
A does not contribute so that only the ΛV component remains of
00
interest. In the static limit Π is simply given by πL (0, k) which in
turn can be identified with the square of the screening scale kS2 in a
medium (Sect. 6.4.1). This implies that the neutrino interacts with the
external electric field as if it had a charge
√
eν = −(CV /e) 2 GF kS2 . (6.102)
In a classical (nondegenerate, nonrelativistic) hydrogen plasma the
screening scale is given by the Debye scale through kS2 = 2kD
2
= 2e2 ne /T
with the
√ electron density ne so that the induced neutrino charge is eν =
CV e 2 2 GF ne /T . This induced charge is explained by the medium
polarization caused by the weak force exerted by the presence of the
neutrino.
While this induced charge is conceptually very interesting it does
not seem to have any immediate practical consequences. Notably, it
is not the relevant quantity for the interaction with a static magnetic
field, i.e. one may not infer that neutrinos move on curved paths in
magnetic fields. The presence of a “neutrino charge” was derived for
the interaction with a static electric field! The relevant form factor
for the interaction with a static magnetic field is identified by noting
that now only the spatial components of Aµ are nonzero. In the static
limit only the component Π00 of the polarization tensor survives when
contracted with Aµ . Then there is no contribution from Λαβ S for the
neutrino interaction with a magnetic field. The contribution from Λαβ A
can be interpreted as a “normal” or “Dirac magnetic moment” induced
by the medium (Semikoz 1987a; D’Olivo, Nieves, and Pal 1989)
√ # ∞ & '
µν = −eCA 2GF 4π dp fe− (p) − fe+ (p) . (6.103)
0
In the limit of a classical plasma this is µν = (eν /2me )(2CA /CV ) where
eν is the induced electric charge of Eq. (6.102).
This induced Dirac magnetic moment is to be compared with the
electron’s Dirac moment e/2me , not with an anomalous moment. The
former arises from the eψ e γµ ψe Aµ coupling, the latter is described by
1
µ ψ σ ψ F µν . This means that the induced dipole moment does
2 e e µν e
not lead to neutrino spin precession—it only couples to left-handed
states. It entails an energy difference between neutrinos moving in
opposite directions along a magnetic field. The transverse part of the
field has no impact on the neutrino—there is no spin precession, and no
curvature of the trajectory. (These conclusions pertain to the limit of
weak magnetic fields. For strong fields the modification of the electron
Particle Dispersion and Decays in Media 241
wavefunctions, i.e. Landau levels rather than plane waves, would have
to be used for a self-consistent treatment of the photon polarization
tensor and thus, for the neutrino coupling to a magnetic field. For a
first discussion see Oraevskiı̆ and Semikoz 1991.)
In summary, on the basis of the existing literature it appears that
the medium-induced electromagnetic form factors of neutrinos are of
practical importance only for the photon decay process that was dis-
cussed in the previous section.
gµν Q2 gµν − Qµ Qν
Dµν (Q) = + + ... (6.104)
m2Z,W m4Z,W
and keep only the first term. (The second term is needed if the con-
tribution of the first one cancels as in a CP symmetric medium—see
38
In the formalism of finite temperature and density (FTD) field theory the am-
plitudes may be written in a more compact form so that the relevant Feynman
graphs reduce to a tadpole and a bubble graph (Nötzold and Raffelt 1988; Nieves
1989; Pal and Pham 1989). Apart from a more compact notation, however, the
FTD formalism leads to the same expressions as the “pedestrian” approach chosen
here.
39
See however Learned and Pakvasa (1995) as well as Domokos and Kovesi-
Domokos (1995) for a discussion of the oscillations of very high-energy cosmic neu-
trinos for which this approximation is not adequate.
242 Chapter 6
(upper sign ν, lower sign ν). Here, nB is the baryon density and
nf − nf
Yf ≡ (6.109)
nB
are the particle number fractions commonly used in astrophysics. Nu-
merically,
√ ρ
2 GF nB = 0.762×10−13 eV (6.110)
g cm−3
with the mass density ρ.
A remark concerning the absolute sign of V is in order. The relative
signs between the different CV ’s can be worked out easily from the
weak interaction structure of the standard model. Also, the relative
sign of the effective neutral-current amplitudes which follow from Z ◦
and W exchange follows directly, for example, from the FTD approach
(Nötzold and Raffelt 1988). Thus to fix the overall sign it is enough
40
In a supernova core or in the early universe it is not possible to distinguish
between a “test neutrino” and a “medium neutrino.” There, one has to study the
nonlinear evolution of the entire ensemble self-consistently (Sect. 9.3.2).
244 Chapter 6
Dicus and Repko (1993) who worked out explicitly the matrix elements
and cross sections. From their results one can extract the forward
scattering amplitude which leads to a refractive index for ν& (- = e, µ, τ )
in a photon bath,
$ ! "%
α GF 4 m2W
nrefr − 1 = 1 + ln 0Eγ 1nγ , (6.114)
4π m2W 3 m2&
ing electric neutrality it is (Botella, Lim, and Marciano 1987; see also
Semikoz 1992 and Horvat 1993)
$ ! " %
3G2F m2τ m2W Yn
Vντ − Vνµ = n B ln −1+ , (6.116)
2π 2 m2τ 3
with a sign change for Vν τ − Vν µ . The shift of the “effective mass” m2eff
is numerically
+ ,2 ρ ω 6.61 + Yn /3
2ω(Vντ − Vνµ ) = 2.06×10−6 eV −3
,
g cm MeV 7
(6.117)
much smaller than the corresponding difference between νe and νµ,τ .
4h2 d3 p2 d3 k
dΓ = P1 · P2 (2π)4 δ 4 (P1 − P2 − K)
2E1 2E2 (2π)3 2ω(2π)3
(6.125)
If one ignores the vacuum mass relative to V one has p1,2 = E1,2 ∓ V
so that to lowest order in V
dΓ V (E1 − E2 )
= αχ . (6.128)
dE2 E12
Therefore, the final-state neutrino spectrum has a triangular shape
where E2 varies between 0 and E1 . The integrated decay rate is
Γ = 12 αχ V (6.129)
Nonstandard Neutrinos
251
252 Chapter 7
95% CL mass limit of 23.8 MeV on the basis of 25 events of the form
τ → 5πντ and 5ππ ◦ ντ where π stands for a charged pion.
Another kinematical method to be discussed in Sect. 11.3.4 uses
the neutrino pulse dispersion from a distant supernova (SN). For νe the
observed neutrinos from SN 1987A gave mνe < ∼ 20 eV, less restrictive
than the tritium experiments. However, if the neutrino pulse from a
future galactic SN will be detected one may be able to probe even a ντ
mass down to the cosmologically interesting 30 eV range (Sect. 11.6)!
For Dirac neutrinos there is another essentially kinematical con-
straint from the SN 1987A neutrino observations. The sterile νDirac
components can be produced in scattering processes by helicity flips.
In a supernova core this effect leads to an anomalous energy drain,
limiting a Dirac mass to be less than a few 10 keV (Sect. 13.8.1).
If one of the neutrinos had a mass near this bound it would be the main
component of the long-sought dark matter of the universe.
Certain scenarios of structure formation currently favor “hot plus
cold dark matter” where neutrinos with mνe + mνµ + mντ ≈ 5 eV play a
sub-dominant dynamical role but help to shape the required spectrum
of primordial density perturbations (Pogosyan and Starobinsky 1995
and references therein). Preferably, the three neutrino masses should
be degenerate rather than one dominating flavor.
If neutrinos were unstable and if they decayed so early that their
decay products were sufficiently redshifted by the expansion of the uni-
verse, the cosmological mass bound can be violated without running
into direct conflict with observations. The excluded range of masses
Nonstandard Neutrinos 259
The expansion rate and thus the energy density of the universe are
well “measured” at the epoch of nucleosynthesis (T ≈ 0.3 MeV) by
the primordial light-element abundances (Yang et al. 1984). This big-
bang nucleosynthesis (BBN) argument has been used to constrain the
number of light neutrino families to Nν < ∼ 3.4 (Yang et al. 1984; Olive
et al. 1990). Even though the measured Z ◦ decay width has established
Nν = 3 (Particle Data Group 1994) the BBN bound remains of interest
as a mass limit because massive neutrinos contribute more than a mass-
less one to the expansion rate at BBN. For a lifetime exceeding about
100 s this argument excludes 500 keV < <
∼ mν ∼ 35 MeV (Kolb et al. 1991;
Dolgov and Rothstein 1993; Kawasaki et al. 1994), with even more re-
strictive limits for Dirac neutrinos (Fuller and Malaney 1991; Enqvist
and Uibo 1993; Dolgov, Kainulainen, and Rothstein 1995). In Fig. 7.2
the region thus excluded is hatched and marked “BBN.”
Kawasaki et al. (1994) have considered the majoron mode ν → ν # χ
(Sect. 15.7) as a specific model for the neutrino decay. Including the
energy density of the scalar χ they find even more restrictive limits
which exclude the region between the dashed lines in Fig. 7.2.
where the unitary matrix U plays the role of the CKM matrix. Unless
otherwise stated ν1 will always refer to the dominant mass admixture
of νe and so forth. It seems plausible that m1 < m2 < m3 , a hierarchy
that is often assumed.
Nonstandard Neutrinos 263
Flavor mixing is the only possibility for members of one family (one
row in Fig. 7.1) to transform into those of a different family. This phe-
nomenon is known as the absence of flavor-changing neutral currents;
it means that the Z ◦ coupling to quarks and leptons, like the photon
coupling, leaves a given superposition of fermions unaltered. For ex-
ample, muons decay only by the flavor-conserving mode µ− → νµ e− ν e
(Fig. 7.4); the experimental upper limits on the branching ratios for
µ− → e− γ and µ− → e− e+ e− are 5×10−11 and 1.0×10−12 , respectively.
In the absence of neutrino masses and mixing the individual lepton fla-
vor numbers are conserved: a lepton can be transformed only into its
partner of the same family, or it can be created or annihilated together
with an antilepton of the same family.
1 G2F
= |Ueh |2 m5 Φ(mh )
τe+ e− 3 (4π)3 h
= |Ueh |2 3.5×10−5 s−1 m5MeV Φ(mh ), (7.9)
The one- and two-photon decay modes arise in the standard model
with mixed neutrinos from the amplitudes shown in Fig. 7.7. Turn first
to the one-photon decay νi → νj γ with the neutrino masses mi > mj .
Nonstandard Neutrinos 265
Fig. 7.7. Feynman graphs for neutrino radiative decays. There are other
similar graphs with the photon lines attached to the intermediate W boson.
where µ2eff ≡ |µij |2 + |-ij |2 , meV ≡ mi /eV, and δm ≡ (m2i − m2j )/m2i .
An explicit evaluation of the one-photon amplitude of Fig. 7.7 yields
for Dirac neutrinos (Pal and Wolfenstein 1982)
- √
µD
ij e 2 GF "
= 2
(m i ± m j ) U&j U&i∗ f (r& ). (7.13)
-D
ij
(4π) &=e,µ,τ
If one inserts the leading term − 32 into the sum in Eq. (7.13) one finds
that its contribution vanishes because the unitarity of U implies that its
rows or columns represent orthogonal vectors. Because the first nonzero
266 Chapter 7
These small numbers imply that neutrino radiative decays are exceed-
ingly slow in the standard model.
Dirac neutrinos would have static or diagonal (i = j) magnetic
dipole moments while the electric dipole moments vanish according to
Eq. (7.13). Their presence would require CP-violating interactions.
Majorana neutrinos, of course, cannot have any diagonal electromag-
netic moments. For µD ii the leading term of Eq. (7.14) in Eq. (7.13)
does not vanish because the unitarity of U implies that the sum equals
unity for i = j. Therefore,
√
µD 6 2 G F me
ii
= mi = 3.20×10−19 meV , (7.16)
µB (4π)2
much larger than the transition moments because it is not GIM sup-
pressed.
The two-photon decay rate νi → νj γγ is of higher order and thus
may be expected to be smaller by a factor of α/4π. However, it is
not GIM suppressed so that it is of interest for a certain range of neu-
trino masses (Nieves 1983; Ghosh 1984). Essentially, the result in-
volves another factor α/4π relative to the one-photon rate, and f (r& )
in Eq. (7.13) is replaced by (mi /m& )2 .
As an example consider the different decay modes for ν3 → ν1 ,
assuming that m3 , m1 and that the mixing angles are small so that
ν3 ≈ ντ and ν1 ≈ νe . Then one has explicitly
Φ(m3 ), ν 3 → ν 1 e+ e− ,
# $
27 α
mτ 4
1 G2 m 5 , ν3 → ν1 γ,
≈ |Ue3 |2 F 33 × 8 4π mW (7.17)
τ 3 (4π)
# $ # $
1 α 2 m3 4
, ν3 → ν1 γγ,
180 4π me
where Φ(mh ) was given in Eq. (7.10) and shown in Fig. 7.6. The γγ
decay dominates in a small range of m3 just below 2me .
Nonstandard Neutrinos 267
Much larger values would obtain with direct r.h. neutrino interac-
tions. For example, in left-right symmetric models there exist heavier
gauge bosons which mediate r.h. interactions; parity violation would
occur because of the mass difference between the l.h. and r.h. gauge
bosons. For a neutrino ν& (flavor + = e, µ or τ ) the dipole moment in
such models is (Kim 1976; Marciano and Sanda 1977; Bég, Marciano,
and Ruderman 1978)
2 + , + ,3
eGF m2W1 3 m2W1
µν = √ 2 m& 1− 2 sin 2ζ + mν"
4
1+ 2 ,
2 2π mW2 m W2
(7.18)
ωP3 "7 8
Γγ = µ2ν with µ2ν = |µij |2 + |-ij |2 , (7.26)
24π i,j
while in the frame of the medium where the photon has the energy ω a
Lorentz factor ωP /ω must be included. The sum includes all final-state
neutrino flavors with mi + ωP ; otherwise phase-space modifications
occur, and even a complete suppression of the decay by a neutrino
mass threshold. In contrast with Eq. (7.25) relevant for the scattering
rate, no destructive interference effects between magnetic and electric
dipole amplitudes occur.
Transition moments would allow for the radiative decay νi → νj γ.
Again, because a neutrino charge radius or anapole moment vanish in
the Q2 → 0 limit relevant for free photons, radiative neutrino decays are
most generally characterized by their magnetic and electric transition
Nonstandard Neutrinos 275
The best limits on µνe are based on the use of reactor neutrinos
as a source. The measurement of the ν e -e scattering cross section by
Reines, Gurr, and Sobel (1976) was interpreted by Kyuldjiev (1984)
to yield a bound of µνe < 1.5×10−10 µB . Since then, the reactor ν e
spectrum has been much better understood. Vogel and Engel (1989)
stressed that a literal interpretation of the old results by Reines, Gurr,
and Sobel (1976) would actually yield evidence for a dipole moment of
about 2−4×10−10 µB . There is, however, a more recent limit of
µν < −12
∼ 3×10 µB . (7.33)
Neutrino Oscillations
8.1 Introduction
If neutrinos do not have novel interactions that allow them to decay
fast then they must obey the cosmological mass limit of mν < ∼ 30 eV.
This is even true for ντ although it could decay sufficiently fast into
the e+ e− νe channel if it had a mass in the 10 MeV range. However, the
absence of γ rays from SN 1987A in conjunction with the neutrino signal
(Sect. 12.5.2) and independently arguments of big-bang nucleosynthesis
(Fig. 7.2) exclude this option. If neutrino masses are indeed so small
then there is no hope for a direct experimental measurement at the
present time, with the possible exception of mνe which could still show
up in tritium β decay or neutrinoless ββ decay experiments as discussed
in Chapter 7.
Pontecorvo (1967) was the first to realize that the existence of sev-
eral neutrino flavors (two were known at the time) allows even very
small masses to become visible.46 A “weak-interaction eigenstate”
which is produced, say, in the neutron decay n → pe− ν e is in gen-
eral expected to be a mixture of neutrino mass eigenstates. The phe-
nomenon of particle mixing (Sect. 7.2.1) is familiar from the quarks
46
Pontecorvo’s (1957, 1958) original discussion referred to ν ↔ ν oscillations in
analogy to the experimentally observed case of K ◦ ↔ K ◦ . For a historical overview
see Pontecorvo (1983). In this book I will not discuss ν ↔ ν oscillations any
further—see Akhmedov, Petcov, and Smirnov (1993) for a recent reexamination of
“Pontecorvo’s original oscillations.”
280
Neutrino Oscillations 281
oscillations, except that here the two helicity components rather than
the flavor components get transformed into each other. Again, this is
a standard effect familiar from the behavior of electrons in magnetic
fields. In astrophysical bodies large magnetic fields exist, especially in
supernovae, so that magnetic helicity oscillations are potentially inter-
esting. However, much larger magnetic dipole moments are required
than are predicted for standard massive neutrinos. Thus, flavor oscil-
lations have rightly received far more attention.
Presently I will develop the theoretical tools for neutrino oscilla-
tions, and summarize the current experimental situation. In Chapter 9
I will discuss the more complicated phenomena that obtain when os-
cillating neutrinos are trapped in a supernova core. The story of solar
neutrinos (Chapter 10) is inextricably intertwined with that of neu-
trino oscillations, especially of the MSW variety. Finally, oscillations
may also play a prominent role for supernova neutrinos and the inter-
pretation of the SN 1987A signal (Chapter 11).
As usual one expands the neutrino fields in plane waves of the form
Ψ(t, x) = Ψk (t) eik·x for which Eq. (8.1) is
Fig. 8.1. Oscillation pattern for two-flavor oscillations (neutrino energy ω).
ρ = 12 (1 + P · σ) and K = ω − b0 + 12 B · σ, (8.18)
∂z P = B × P or ∂t P = B × P. (8.19)
Eν 1 eV2
&osc = 2.48 m . (8.24)
1 MeV ∆m2
Therefore, the modulation of the flavor content of a neutrino beam can
occur on large, even astronomical length scales.
There exists a large number of oscillation searches using terrestrial
neutrino sources (reactors, accelerators); for detailed references see Par-
ticle Data Group (1994). In Fig. 8.5 (curves a–f ) I show the most
restrictive limits on oscillations between the known neutrinos where
Fig. 8.5. Experimental limits on neutrino masses and mixing angles. Re-
actors, νe disappearance: (a) Bugey 4 (Achkar et al. 1995), superseding the
Gösgen limits (Zacek et al. 1986); (b) Kurchatov Institute (Vidyakin et al.
1987, 1990, 1991). Accelerator experiments: (c) BNL Experiment 776, wide-
band beam, νe and ν e appearance (Borodovsky et al. 1992). (d) BNL Ex-
periment 734, measurement of νe /νµ ratio (Ahrens et al. 1985). (e) Fermilab
Experiment 531, ντ appearance (Ushida et al. 1986); similar constraints were
reported by the CHARM II Collaboration (1993). (f) CDHS Experiment,
νµ disappearance (Dydak et al. 1984). (g) Anticipated range of sensitivity
for the CHORUS and NOMAD experiments which are currently taking data
at CERN (DiLella 1993; Winter 1995).
290 Chapter 8
the analysis was always based on the assumption that two-flavor os-
cillations dominate. The disappearance experiments, of course, also
constrain oscillations into hypothetical sterile neutrinos.
Even though the experimental results look very impressive, a glance
on the CKM matrix Eq. (7.6) reveals that one could not yet have ex-
pected to see oscillations in the νe ↔ ντ or νµ ↔ ντ channel if the
neutrino mixing angles are comparably small. It is very encouraging
that the NOMAD and CHORUS experiments which are currently tak-
ing data at CERN (DiLella 1993; Winter 1995) anticipate a range of
sensitivity (curve g in Fig. 8.5) which is promising both in view of the
possible cosmological role of a mν in the 10 eV range and the small
mixing angles probed. Other future but less advanced projects for ter-
restrial oscillation searches were reviewed by Schneps (1993, 1995).
At the time of this writing the LSND Collaboration has reported a
signature that is consistent with the occurrence of ν µ → ν e oscillations
(Athanassopoulos et al. 1995). If this interpretation is correct, the
corresponding ∆m2 would exceed about 1 eV2 , while sin2 2θ would be a
few 10−3 . The status of this claim is controversial at the present time—
see, e.g. Hill (1995). No doubt more data need to be taken before one
can seriously begin to believe that neutrino oscillations have indeed
been observed.
p + A → n + π/K + . . .
π/K → µ+ (µ− ) + νµ (ν µ )
µ+ (µ− ) → e+ (e− ) + νe (ν e ) + ν µ (νµ ).
(8.25)
Fig. 8.6. Limits on neutrino masses and mixing angles from atmospheric
neutrinos. (a) The shaded area is the range of masses and mixing angles
required to explain the νe /νµ anomaly at Kamiokande (Fukuda et al. 1994);
the star marks the best-fit value for the mixing parameters. The hatched
areas are excluded by: (b) νe /νµ ratio at Fréjus (Fréjus Collaboration 1990,
1995; Daum 1994). (c) Absolute rate and (d) stopping fraction of upward go-
ing muons at IMB (Becker-Szendy et al. 1992). Also shown are the excluded
areas from the experimental limits of Fig. 8.5.
and 13000 km are available.48 The energy spectrum and absolute nor-
malization of the flux must be determined by calculations and thus is
probably uncertain to within about ±30% while the νe /νµ flavor ratio
is likely known to within, say, ±5%.
Several underground proton decay experiments have reported mea-
surements of atmospheric neutrinos. The Fréjus detector (an iron
calorimeter) saw the expected νe /νµ flavor ratio and thereby excluded
the range of masses and mixing angles marked b in Fig. 8.6 for νe -νµ
and νµ -ντ oscillations (Fréjus Collaboration 1990, 1995; Daum 1994).
Instead of measuring the neutrinos directly one may also study the
flux of secondary muons produced by interactions in the rock surround-
48
The effect of matter must be included for νe -νµ atmospheric neutrino oscilla-
tions. For a recent detailed analysis see Akhmedov, Lipari, and Lusignoli (1993).
292 Chapter 8
ing the detector. (Electrons from νe interactions range out much faster
in the rock and so one expects mostly muons from νµ ’s.) This method is
sensitive to the high-energy spectral regime of the atmospheric νµ flux.
Moreover, one may select upward going muons which are produced from
νµ ’s which traversed the entire Earth and thus have a large oscillation
length available. The IMB detector excludes range c by this method
(Becker-Szendy et al. 1992). Also, one may determine the fraction of
muons stopped within the detector to those which exit, allowing one to
constrain a spectral deformation caused by the energy dependence of
the oscillation length. Range d is excluded by this method according
to the IMB detector (Becker-Szendy et al. 1992).
However, several detectors see a substantial deficit of atmospheric
νµ ’s relative to νe ’s, a finding usually expressed in terms of a “ratio of
ratios,” i.e. the measured over the expected ratio of e-like over µ-like
events (Fig. 8.7). While this procedure is justified because it is largely
free of the uncertain absolute flux normalization, one must be careful
at interpreting the significance of the flux deficit. The error of a mea-
sured ratio does not follow a Gaussian distribution; a representation
like Fig. 8.7 tends to overemphasize the significance of the discrepancy
(Fogli and Lisi 1995).
Fig. 8.7. Measured ratio of the atmospheric νµ /νe fluxes relative to the
expected value (“ratio of ratios”) in five detectors. Where two results are
shown they refer to different signatures or data samples. (See Goodman
1995 for references.)
Neutrino Oscillations 293
This equation defines implicitly the mixing angle θ as well as the os-
cillation length &osc in the medium in terms of the masses, the vacuum
mixing angle θ0 , and the electron density ne .
Neutrino Oscillations 295
Fig. 8.8. Mixing angle, oscillation length, and neutrino dispersion relation
as a function of the electron density. The medium was taken to have equal
numbers of protons and neutrons (Ye = 12 ), the ratio of neutrino masses was
taken to be m1 : m2 = 1 : 2, and sin2 2θ0 = 0.15.
296 Chapter 8
Explicitly one finds for the mixing angle and the oscillation length
in the medium the following expressions,
sin 2θ0
tan 2θ = ,
cos 2θ0 − ξ
sin 2θ0
sin 2θ = 2 , (8.30)
[sin 2θ0 + (cos 2θ0 − ξ)2 ]1/2
where
√
2 GF ne 2ω −7 Ye ρ ω eV2
ξ≡ = 1.53×10 , (8.31)
m22 − m21 g cm−3 MeV m22 − m21
and
4π ω sin 2θ
&osc = . (8.32)
m22 2
− m1 sin 2θ0
For m2 > m1 these functions are shown in Fig. 8.8; they exhibit a
“resonance” for cos 2θ0 = ξ.
The dispersion relation has two branches which in vacuum corre-
spond to ω1,2 = (k 2 − m21,2 )1/2 . In the relativistic limit they are
m21 + m22 √
ω1,2 − k = + 2 GF nB (Ye − 12 )
4k
m2 − m21 . 2 /1/2
± 2 sin 2θ0 + (cos 2θ0 − ξ)2 , (8.33)
4k
a result schematically shown in Fig. 8.8. The “resonance” of the mix-
ing angle corresponds to the crossing point of the two branches of the
dispersion relation. Of course, the levels do not truly cross, but rather
show the usual “repulsion.”
Because the medium effect changes sign for antineutrinos, a reso-
nance occurs between νe and νµ if m2 > m1 , while none occurs between
ν e and ν µ . If the mass hierarchy is the other way round, a resonance
occurs for ν e and ν µ , but not for νe and νµ .
when the quantity ξ of Eq. (8.31) is much smaller than unity. In the
opposite limit one finds for the mixing angle in the medium
m2 − m21
sin 2θ = √ 2 sin 2θ0 . (8.34)
2 GF ne 2ω
The oscillation length becomes
2π g cm−3
&osc = √ = 1.63×104 km , (8.35)
2 GF ne Ye ρ
independent of the neutrino masses or energy.
Typically the effect of the medium is, therefore, to suppress the
mixing angle and thus the possibility to observe oscillations. Of course,
for normal materials with a density of a few g cm−3 and neutrino masses
in the eV range one needs TeV neutrino energies for the medium to be
relevant at all. For a review of the impact of oscillations in a medium
on neutrino experiments or the observation of atmospheric, solar, or
supernova neutrinos see, for example, Kuo and Pantaleone (1989).
This translates into ∇θ = 12 ξ (sin2 2θ/ sin 2θ0 ) ∇ ln ne while &osc is given
by Eq. (8.32) and ξ by Eq. (8.31). Thus, the adiabatic condition is
the other (Fig. 8.8, lowest panel) when it moves across the resonant
density region.
A linear density profile near the resonance region, which is always
a first approximation, yields the “Landau-Zener probability”
p = e−πγ/2 (8.41)
which was first derived in 1932 for atomic level crossings. The adia-
baticity parameter γ was defined in Eq. (8.39); in the adiabatic limit
γ ' 1 one recovers p = 0.
For a variety of other density profiles and without the assumption
of a small mixing angle one finds a result of the form
!
e−(πγ/2) F − e−(πγ/2) F
p= , (8.42)
1 − e−(πγ/2) F !
where F # = F/ sin2 θ0 and F is an expression characteristic for a given
density profile. For a linear profile ne ∝ r one has F = 1 so that for
a small mixing angle one recovers the Landau-Zener probability. The
profile ne ∝ r−1 leads to F = cos2 2θ0 / cos2 θ0 . Of particular interest is
the exponential ne ∝ e−r/R0 which yields
F = 1 − tan2 θ0 . (8.43)
All of these results are quoted after the review by Kuo and Pantaleone
(1989) where other special cases and references to the original literature
can be found.
Going beyond the Landau-Zener approximation requires assuming
one of the above specific forms for ne (r) for which analytic results ex-
ist. Recently, Guzzo, Bellandi, and Aquino (1994) used a somewhat
different approach which is free of this limitation. They derived an
approximate solution to the equivalent of Eq. (8.36) by the method of
stationary phases for the space-ordered exponential. They found
' (2
1 − γ# 2γ # . /
2 2 2
p= sin (θ0 − θ) + cos (θ 0 − θ) + cos (θ0 + θ) ,
1 + γ# (1 + γ # )2
(8.44)
∆m2 MeV
γ= 1
× 103 sin 2θ0 tan 2θ0 . (8.45)
6
meV2 Eν
An electron density at the center of nc = 1.6×1026 cm−3 yields
meV2 Eν
ξc = 40 . (8.46)
∆m2 MeV
The quantity ξ was defined in Eq. (8.31).
It is further assumed that all neutrinos are produced with a fixed
energy by a point-like source at the solar center. They will encounter
a resonance on their way out if ξc > cos 2θ0 . In this case one may
calculate the “jump probability” p according to Eq. (8.42) with F from
Eq. (8.43) for the exponential profile. The survival probability is then
given by Eq. (8.40) with the mixing angle at the solar center
cos 2θ0 − ξc
cos 2θ = . (8.47)
[(cos 2θ0 − ξc )2 + sin2 2θ0 ]1/2
where they are not. The dotted line marks ξc = cos 2θ0 and thus indi-
cates for which neutrino parameters a resonance occurs on the way out
of the Sun.
Fig. 8.9. The MSW triangle for the simplified solar model discussed in the
text with Eν = 1 MeV.
Fig. 8.10. The MSW bathtub for the simplified solar model discussed in the
text with ∆m2 = 3×10−5 eV2 and sin2 2θ0 = 0.01.
Neutrino Oscillations 303
Oscillations
of Trapped Neutrinos
9.1 Introduction
So far oscillations were discussed in the context of “beam experiments”
where neutrinos of a known flavor are produced at a certain location,
for example in a power reactor or in the Sun, then propagate over a
distance, and are then detected with a device that allows one to dis-
tinguish between different flavors. It was assumed that there were no
interactions between the production and detection point, with the pos-
sible exception of decays, allowing one to treat neutrino oscillations as
a simple propagation phenomenon fully analogous to light propagation
in an optically active medium.
This approach is not adequate for the early universe and a young SN
core. In both cases there are frequent neutrino collisions which affect
the free evolution of the phases. The impact of these collisions can
be understood if one assumes for the purpose of illustration that one
310
Oscillations of Trapped Neutrinos 311
neutrino flavor (say νe ) scatters with a rate Γ while the other (say νµ )
does not. An initial νe will begin to oscillate into νµ . The probability
for finding it in one of the two flavors evolves as previously discussed
and as shown in Fig. 9.1 (dotted line). However, in each collision the
momentum of the νe component of the superposition is changed, while
the νµ component remains unaffected. Thus, after the collision the two
flavors are no longer in the same momentum state and so they can no
longer interfere: each of them begins to evolve separately. This allows
the remaining νe to develop a new coherent νµ component which is made
incoherent in the next collision, and so forth. This process will come
into equilibrium only when there are equal numbers of νe ’s and νµ ’s.
This decoherence effect is even more obvious when one includes the
possibility of νe absorption and production by charged-current reactions
νe n ↔ pe. Because of oscillations an initial νe is subsequently found
to be a νµ with an average probability of 12 sin2 2θ (mixing angle θ)
and as such cannot be absorbed, or only by the reaction νµ n ↔ pµ
if it has enough energy. The continuous emission and absorption of
νe ’s spins off a νµ with an average probability of 12 sin2 2θ in each
collision! Chemical relaxation of the neutrino flavors will occur with
an approximate rate 12 sin2 2θ Γ where Γ is a typical weak interaction
rate for the ambient physical conditions. An initial νe population
turns into an equal mixture of νe ’s and νµ ’s as shown schematically
in Fig. 9.1 (solid line).
Fig. 9.1. Neutrino oscillations with collisions (solid line). In the absence
of collisions and for a single momentum one obtains periodic oscillations
(dotted line), while for a mixture of energies the oscillations are washed out
by “dephasing” (dashed line).
312 Chapter 9
Ṗ = V × P − DPT . (9.1)
The first part is the previous precession formula, except that here a
temporal evolution is appropriate since one has in mind the evolution
of a spatially homogeneous ensemble rather than the spatial pattern of
a stationary beam. The “magnetic field” V is
sin 2θ
2π
V= 0 , (9.2)
tosc
cos 2θ
where θ is the mixing angle in the medium and tosc the oscillation
period. They are given in terms of the neutrino masses and momen-
tum, the vacuum mixing angle, and the medium density by Eqs. (8.30)
and (8.32) where strictly speaking the neutrino energy is to be replaced
by its momentum as we have turned to temporal rather than spatial
oscillations. The damping parameter D is determined by the scattering
amplitudes on the background.
The evolution described by Eq. (9.1) is a precession around the
“magnetic field” V, combined with a shrinking of the length of P to
zero. This final state corresponds to ρ = 12 where both flavors are
equally populated, and with vanishing coherence between them. For
1
D = t−1osc and sin 2θ = 2 the evolution of the flavors is shown in Fig. 9.1
(solid line). Ignoring the wiggles in this curve it is an exponential as
can be seen by multiplying both sides of Eq. (9.1) with P/P 2 which
Oscillations of Trapped Neutrinos 315
(Dolgov 1981)
' ( ' (
ρij (p) = a†j (p)ai (p) and ρij (p) = b†i (p)bj (p) , (9.3)
where ai (p) and a†i (p) are the destruction and creation operators for
neutrinos of flavor i in mode p while b is for antineutrinos which other-
wise are referred to by overbarred quantities. The reversed order of the
flavor indices in the definition of ρ(p) guarantees that both matrices
transform in the same way under a unitary transformation in flavor
space. Also, for brevity $ | . . . | % is always written as $. . .%.
The diagonal elements of ρp and ρp are the usual occupation num-
bers while the off-diagonal ones represent relative phase information.
In the nondegenerate limit, up to a normalization ρp plays the role of
the previously defined single-particle density matrix. Therefore, the
ρp ’s and ρp ’s are well suited to account simultaneously for oscillations
and collisions. In fact, one can argue that a homogeneous neutrino
ensemble is completely characterized by these “matrices of densities”
(Sigl and Raffelt 1993). It remains to derive an equation of motion
which in the appropriate limits should reduce to the previous preces-
sion equation, to a Boltzmann collision equation, and to Stodolsky’s
damping equation (9.1), respectively.
ρ̂ij (p, t) ≡ a†j (p, t)ai (p, t) and ρ̂ij (p, t) ≡ b†i (p, t)bj (p, t) . (9.7)
and similar for ρ̂p . With H = H0 from Eq. (9.6) one finds
i∂t ρ̂p = [Ω0p , ρ̂p ] and i∂t ρ̂p = −[Ω0p , ρ̂p ] . (9.9)
and an analogous equation for ρp (t). These equations are exact, but
they are not a closed set of differential equations for the ρp and ρp . To
this end one needs to perform a perturbative expansion.
To first order one may set the interacting fields B(t) and Ψ(t) on the
r.h.s. of Eq. (9.10) equal to the free fields50 B0 (t) and Ψ0 (t). Under the
assumption that the original state contained no correlations between
the neutrinos and the background the expectation value factorizes into
a medium part and a neutrino part. With Wick’s theorem and ignoring
fast-varying terms such as b† b† it can be reduced to an expression which
contains only ρp ’s and ρp ’s. The result gives the forward-scattering or
refractive effect of the interaction.
To include nonforward collisions one needs to go to second order
in the perturbation expansion. At a given time t a general operator
ξ(t) = ξ(B(t), Ψ(t)) which is a functional of B and Ψ is to first order
) t * +
ξ(t) = ξ0 (t) + i dt" Hint
0
(t − t" ), ξ0 (t) , (9.11)
0
0
where ξ0 and Hint are functionals of the freely evolving fields B0 (t)
and Ψ0 (t). Applying this general iteration formula to the operator
ξ = [Hint (B, Ψ), ρ̂p ] which appears on the r.h.s. of Eq. (9.10) one
arrives at
* + '* +(
ρ̇p (t) = −i Ω0p , ρp (t) + i 0
Hint (t), ρ̂0p
) t '* * ++(
− dt" 0
Hint (t − t" ), Hint
0
(t), ρ̂0p , (9.12)
0
and similar for ρp (t). The second term on the r.h.s. is the first-order
refractive part associated with forward scattering. The second-order
term contains both forward- as well as nonforward-scattering effects.
50
These free operators are the solutions of the equations of motion in the absence
of Hint . However, internal interactions of the medium such as nucleon-nucleon
scattering are not excluded. Moreover, Ψ(0) = Ψ0 (0) etc. are taken as initial
conditions for the interacting fields. Also, the mass term is ignored in the definition
of Ψ0 ; its effect is included only in the first term on the r.h.s. of Eq. (9.10), the
“vacuum oscillation term.” Therefore, the free creation and annihilation operators
vary as a0j (p, t) = aj (p, 0)e−ipt etc. for all flavors with p = |p|. This implies that
the operators ρ̂0 (p) and ρ̂0 (p), which are constructed from the free a’s and b’s, are
time independent.
Oscillations of Trapped Neutrinos 319
Because all operators on the r.h.s. of Eq. (9.12) are free the expec-
tation values in the first- and second-order term factorize between the
neutrinos and the medium. This leads one to equations for ρ̇p (t) and
ρ̇p (t) which on the r.h.s. involve only ρp (t) and ρp (t) as well as $ρ̂0p %
and $ρ̂0p % besides expectation values of B operators.
The interactions described by Hint are taken as individual, isolated
collisions where the neutrinos go from free states to free states as in
ordinary scattering theory. The duration of one collision (the inverse
of a typical energy transfer) is assumed to be small relative to the
time scale over which the density matrices vary substantially, i.e. small
relative to the oscillation time and the inverse collision frequency. Phys-
ically this amounts to the restriction that the neutrino collision rate is
small enough that multiple-scattering effects can be ignored. Further,
it is assumed that the medium is not changed much by the interactions
with the neutrino ensemble, allowing one to neglect evolution equations
for the medium variables which can thus be taken to be externally pre-
scribed, usually by conditions of thermal equilibrium. If the medium
is not stationary it is assumed that the time scale of variation is large
compared to the duration of typical neutrino-medium collisions.
One may then choose the time step of iteration t in Eq. (9.12)
both small relative to the evolution time scale and large relative to the
duration of one collision. Under these circumstances the time integral
can be extended to infinity while setting ρp (t) equal to ρp (0) = $ρ̂0p %.
This leads to
* + '* +(
ρ̇p (0) = −i Ω0p , ρp (0) + i 0
Hint (0), ρ̂0p
) +∞ '* * ++(
0 0
− 1
2
dt Hint (t), Hint (0), ρ̂0p , (9.13)
−∞
/ /
and similar for ρp . Here, 0∞ dt $. . .% was replaced by 12 −∞ +∞
dt $. . .%.
The difference between these expressions corresponds to a principle-
part integral which leads to a second-order correction to the refractive
term which is ignored. In the form of Eq. (9.13) the time integral leads
to energy conservation in individual collisions.
An explicit evaluation of the r.h.s. of Eq. (9.13) for a given interac-
tion model yields the desired set of differential equations for the ρp ’s
and ρp ’s at time t = 0. It will be valid at all times if the correlations
built up by neutrino collisions are “forgotten” before the next colli-
sion occurs. This assumption corresponds to “molecular chaos” in the
derivation of the usual Boltzmann equation.
320 Chapter 9
where na ≡ $Baµ %Pµ /P0 where P is the neutrino four momentum and
thus P/P0 their four velocity. In an isotropic medium the spatial parts
of $Baµ % vanish so that na is the number density of fermions a. If the
medium is unpolarized, axial currents do not contribute.
In the standard model with the coupling constants of Appendix B
one finds for an isotropic, unpolarized medium of protons, neutrons,
and electrons,
* √ +
iρ̇p = (Ω0p + 2 GF N# ), ρp ,
* √ +
−iρ̇p = (Ω0p − 2 GF N# ), ρp , (9.18)
neutrinos are important only in young SN cores where one may ignore
antineutrinos. With the total neutrino matrix of densities
)
ρ≡ dp ρp , (9.20)
where the energy transfer ∆0 can be both positive and negative. In the
ultrarelativistic limit the neutrino tensor can be written as
N µν = 12 (U µ U "ν + U "µ U ν − U · U " g µν − i,µναβ Uα Uβ" ), (9.25)
where U ≡ K/K 0 and U " ≡ K " /K 0 are the neutrino four velocities.
Therefore, N µν is an even function of K and K " . Note that the defini-
tion Eq. (9.25) differs slightly from the corresponding Eq. (4.17).
The first two terms of the collision integral Eq. (9.22) are due to
neutrino scattering off the medium. The positive term represents gains
from scatterings νp! → νp while the negative one is from losses by the
inverse reaction. The third and fourth expressions account for pair
processes, i.e. the creation or absorption of νp ν p! by the medium. The
pair terms are found by direct calculation or from the scattering ones
by “crossing,”
P → −P and ρp → (1 − ρp ) . (9.26)
For example, the reaction νp X → X " νp! transforms to X → X " ν p νp!
under this operation where X and X " represent medium configurations.
The collision integral for ρp is found by direct calculation or by
applying the crossing operation Eq. (9.26) to all neutrinos and antineu-
trinos appearing in Eq. (9.22). The neutrino gain terms then transform
to the antineutrino loss terms and vice versa.
324 Chapter 9
Ṗp,coll = − 12 Γp G × (G × Pp ). (9.30)
Thus one naturally recovers Stodolsky’s damping term Eq. (9.1) with
D = 12 Γp |G|2 . For νe and νµ and if one writes G = diag(gνe , gνµ ) in
the weak interaction basis D = 12 Γp (gνe − gνµ )2 . This representation
reflects that the damping of neutrino oscillations depends on the dif-
ference of the scattering amplitudes: D is the square of the amplitude
difference, not the difference of the squares. If one flavor does not
scatter at all, D is half the scattering rate of the active flavor.
Collisions thus lead to chemical equilibrium as discussed in the in-
troduction to this chapter and as shown in Fig. 9.1. However, the
simple exponential damping represented by Stodolsky’s formula can be
reproduced only in the limit of vanishing energy transfers in collisions,
an assumption which amounts to separating the neutrino momentum
degrees of freedom from the flavor ones. In a more general case the evo-
lution is more complicated. In particular, collisions usually lead to a
transient flavor polarization in an originally unpolarized ensemble if the
momentum degrees of freedom were out of equilibrium. Still, the neu-
trinos always move toward kinetic and chemical equilibrium under the
action of the collision integral Eq. (9.22) in the sense that the properly
defined free energy never increases (Sigl and Raffelt 1993).
Here,
If in a collision |p| = |p" | and thus sp = sp! one recovers the previous
result ṅνx = 0. However, if the mixing angle is a function of the neutrino
momentum, NC collisions do lead to flavor conversion and thus to the
damping of oscillations.
Of course, if only true NC interactions existed, the mixing angle
in the medium would be fixed at its vacuum value and so no flavor
Oscillations of Trapped Neutrinos 329
where ψp , ψn , and ψe are the proton, neutron, and electron Dirac fields,
respectively, while CV = 1 and CA = 1.26 are the dimensionless CC
vector and axial-vector nucleon coupling constants.
These expressions are defined for both positive and negative energy
#
transfer ∆0 because P−P plays the role of an absorption rate for an-
tineutrinos with physical (P0 > 0) four momentum while A#−P plays
that of a production rate. Put another way, A#∆ and P∆ #
represent the
rate of absorption or production of lepton number of type ., indepen-
dently of the sign of ∆0 .
It is useful to define a flavor matrix of production rates which in
the weak basis has the form
Pe 0 0
1 ∆ µ
P∆ ≡ 0 P∆ 0 . (9.46)
2 τ
0 0 P∆
An analogous definition pertains to A∆ . Then one finds for the CC
collision integrals (Sigl and Raffelt 1993)
ρ̇p,CC = {PP , (1 − ρp )} − {AP , ρp } ,
ρ̇p,CC = {A−P , (1 − ρp )} − {P−P , ρp } , (9.47)
where {·, ·} is an anticommutator. The kinetic term for ρp is related to
that for ρp by the crossing relation Eq. (9.26).
The r.h.s. of Eq. (9.47) for ρp is the difference between a gain and a
loss term corresponding to the production or absorption of a νp . For a
single flavor they take on the familiar form PP (1 − fp ) and AP fp where
(1 − fp ) is the usual Pauli blocking factor.
of freedom, and one may use the weak-damping limit where neutrino
oscillations are much faster than their rates of collision or absorption.
Then one finds for two flavors (Raffelt and Sigl 1993)
where fpe and fpx are the occupation numbers for νe and νx as in
Sect. 9.3.5. The corresponding equation for νe is found by exchang-
ing e ↔ x everywhere.
For νx = νµ flavor conversion can build up a nonvanishing muon den-
sity in a SN core because they are light enough to be produced initially
when the electron chemical potential is on the order of 200−300 MeV.
For νx = ντ or some sterile flavor, the direct production or absorption
is not possible, AxP = PPx = 0. This simplifies Eq. (9.48) considerably,
* +
f˙px = 14 s2p (2 − fpx − fpe ) PPe − (fpx + fpe ) AeP . (9.49)
where fpe and fpx are the occupation numbers of νe and νx which are
given by Fermi-Dirac distributions because kinetic equilibrium was as-
sumed for both flavors. Also, 14 s2p = 14 sin2 2θE = θE2 for small mixing
angles.
A summation over different species a of medium fermions was re-
stored; gea and gxa are dimensionless effective NC coupling constants of
νe and νx to fermion species a. For a = n or p these constants are
the same for all active neutrino species. Electrons as scattering targets
are very relativistic so that they may be classified into a l.h. and a
Oscillations of Trapped Neutrinos 333
where ρ14 is the density in units of 1014 g cm−3 and as usual Yj gives the
abundance of species j relative to baryons. The approximate parameter
334 Chapter 9
For l.h. electrons the coupling strengths are different, and they are
relativistic so that recoil effects are not small. However, they are degen-
erate so that their contribution is expected to be smaller than the ep
process. It is not entirely negligible, however, especially if the nucleon
contribution is partly suppressed by many-body effects. The transition
rates were worked out in detail by Raffelt and Sigl (1993) for entirely
degenerate leptons. They found
Fp = (1 − η 5 )/(1 − η 3 ),
Fe = ( 56 + 12 η + 14 η 2 + 1 3
12
η ) (1 − η)3 /(1 − η 3 ), (9.66)
336 Chapter 9
Fig. 9.2. Dependence of the flavor relaxation rate on η = µνx /µνe for the
case of a “large” ∆m2 Eq. (9.66) and a “small” ∆m2 Eq. (9.69). Fe is
the contribution of electron targets (νe → eν) while Fp is from protons
(ep ↔ nνe ).
and numerically
5 65/3 5 62
3G2F Yp ρ µνe
np µ2νe = 1.0×109 s−1 , (9.67)
5π 10 g cm−3
14 µe
where Fp = 1 and
9 −1 137
5
η − 40
+ 34 (1 + η 4 ) log η + 53 η + η 3 − 101 4
120
η − 15 η 5
Fe = 8 (9.69)
3
(1 − η 3 )
Fig. 9.3. Contour plot for log(τ θ02 ) with τ in seconds according to Eqs. (9.64)
and (9.68), taking Fe = 0, Fp = 1, CV2 +3CA 2 = 4, and µ /µ = 1. (Adapted
νe e
from Raffelt and Sigl 1993.)
Because the νx are assumed to escape freely one may set fpx = 0 on
the r.h.s. of these equations. Moreover, for the νx coupling constants
one may use gx = 0 because it is sterile. For an isotropic medium
and taking an angular average of the scattering rate as in the previous
section one finds
5 ) 6
.
f˙px = 1 2
s
4 p
PEe + (gea )2 dp "
WEa ! E fpe! . (9.72)
a
the volume loss rates for lepton number and energy can be easily de-
termined.
Turn first to the case ∆m2 > 2
∼ (100 keV) so that one may use the
vacuum mixing angle in a SN core. One must now include both the CC
process ep → nνx as well as the NC process νe N → N νx which is not
suppressed because for a sterile νx there is no destructive interference
effect. For nondegenerate nucleons and using CV2 + 3CA2 ≈ 4 for both
CC and NC processes one easily finds from the results of Sect. 9.5.2
2G2F
ṅL = − 14 sin2 2θ0 nB (Yp + 14 ) µ5νe , (9.75)
5π 3
and the same for Q̇ with 15 µ5νe → 16 µ6νe . Electrons as NC scattering
targets may be neglected because of their degeneracy. Lepton number
is thus lost at a rate
where
5/3 5/3
τ −1 = 35 (36π)1/3 G2F nB = 7.7×1010 s−1 ρ15 . (9.77)
Because Yνe is about YL /4, lepton number and energy are lost at
about a rate of sin2 2θ0 1010 s−1 . Because the SN 1987A signal lasted
for several seconds, a conflict with these observations is avoided if
sin2 2θ0 <
∼ 10
−10
(9.78)
(Kainulainen, Maalampi, and Peltoniemi 1991; Raffelt and Sigl 1993).
If mνx > ∼ 1 MeV the assumed mixing with νe allows for decays νx →
νe e e and νx → νe e− e+ γ. The resulting γ signal from the SN 1987A
− +
Solar Neutrinos
10.1 Introduction
The Sun, like other hydrogen-burning stars, liberates nuclear binding
energy by the fusion reaction
341
342 Chapter 10
Fig. 10.1. Solar neutrino flux at Earth according to the Bahcall and
Pinsonneault (1995) solar model. Upper panel: Continuum spectra in
cm−2 s−1 MeV−1 , line spectra in cm−2 s−1 . Solid lines are the sources of
dominating experimental significance. Above: Range of sensitivity of cur-
rent and near-future solar neutrino experiments. Lower panel: Cumulative
spectrum integrated from a given energy to infinity.
tional sensitivity it is a true “neutrino telescope” and for the first time
established that indeed neutrinos are coming from the direction of the
Sun. It is perplexing, however, that the measured flux is less suppressed
relative to solar-model predictions than that found in the Homestake
experiment. This is the reverse from what would be expected on the
grounds that 37 Cl is sensitive to boron and beryllium neutrinos.
The situation changed yet again when the experiments SAGE (So-
viet-American Gallium Experiment) and GALLEX began to produce
data in 1990 and 1991, respectively. They are radiochemical experi-
ments using the reaction νe + 71 Ga → 71 Ge + e− which has a threshold
of 233 keV. Therefore, these experiments pick up the dominant pp neu-
trino flux which can be calculated from solar models with a precision of
a few percent unless something is radically wrong with our understand-
ing of the Sun. Therefore, a substantial deficit of measured pp neutrinos
would have been a “smoking gun” for the occurrence of neutrino os-
cillations. While the first few exposures of SAGE seemed to indicate
a low flux, the good statistical significance of the data that have since
been accumulated by both experiments indicate a flux which is high
enough so that no pp neutrinos are reported missing, but low enough
to confirm the existence of a significant problem with the high-energy
part of the spectrum (beryllium and boron neutrinos).
With four experiments reporting data, which represent three differ-
ent spectral responses to the solar neutrino flux, the current attention
has largely shifted from a comparison between experiments and theoret-
ical flux predictions to a “model-independent analysis” which is based
on the small number of possible source reactions each of which produces
neutrinos of a well-defined spectral shape. This sort of analysis cur-
rently indicates a lack of consistency among the experiments which can
be brought to perfect agreement if neutrinos are assumed to oscillate.
Even though the attention has currently shifted away from theo-
retical solar neutrino flux predictions it should be noted that in recent
years there has been much progress in a quantitative theoretical treat-
ment of the Sun. Independently of the interest in the Sun as a neutrino
source it serves as a laboratory to test the theory of stellar structure
and evolution. Particularly striking advances have been made in the
field of helioseismology. There are two basic vibration patterns for the
Sun, one where gravity represents the restoring force (g-modes), and
normal “sound” or pressure (p) modes. The former are evanescent in
the solar convection zone (depth about 0.3 R" from the surface) and
have never been unambiguously observed. The oscillation period of the
highest-frequency g-modes would be about 1 h.
Solar Neutrinos 345
Fig. 10.2. Reaction chains PPI−PPIII and CNO tri-cycle. Nuclear reactions,
including decays, are marked with a bullet (•). Average (av) and maximum
(max) energies in MeV are given for the neutrinos. For photons (wavy
arrows) numbers in brackets refer to the total energy of a cascade; otherwise
it is the energy of a monochromatic γ line.
Solar Neutrinos 347
nuclei, the usual weak decay by e+ νe emission is not possible and so the
conversion proceeds by electron capture, leading to the emission of an
almost monochromatic neutrino. In about 10% of all cases the capture
reaction goes to the first excited state (478 keV) of 7 Li so that there are
two neutrino lines.
Instead of an electron, 7 Li very rarely captures a proton and forms
8
B which subsequently decays into 8 Be, a nucleus unstable against spon-
taneous fission into 4 He + 4 He. This PPIII termination occurs in about
0.02% of all cases, too rare to be of importance for nuclear energy gen-
eration. Its importance arises entirely from the high energy of the 8 B
neutrinos which are the ones least difficult to measure.
The neutrino-producing reactions are summarized in Tab. 10.1 with
their maximum energies, and with the resulting neutrino flux at Earth
found in the solar model of Bahcall and Pinsonneault (1995).
Apart from small screening and thermal broadening effects, the
spectral shape for each individual source is independent of details of
the solar model. The pp, 13 N, 15 O, and 17 F reactions are allowed or
superallowed weak transitions so that their spectra are
! "1/2
dN/dEν = A (Q + me − Eν ) (Q + me − Eν )2 − m2e Eν2 F ,
(10.2)
where Q is the maximum e+ kinetic energy and also the maximum
νe energy, A is a normalization constant, and F is a function of Ee+
which takes the e+ final-state interactions into account. For the low-Z
nuclei under consideration this correction is small for most of the neu-
trino spectrum. With F = 1 the normalization constants are given in
Tab. 10.2. In Fig. 10.3 the normalized spectra from the pp and the 15 O
processes are shown where the pp spectrum was taken from the tabula-
tion of Bahcall and Ulrich (1988). Using Eq. (10.2) instead would cause
a change so small that it would be nearly hidden by the line width of
the curve in Fig. 10.3.
where the quality of the fit is equally good as that for the 8 B neutrinos.
350 Chapter 10
Fig. 10.4. Normalized spectra of the thermally and Doppler broadened beryl-
lium neutrino lines where E0 is the corresponding laboratory energy (Bahcall
1994). The line shapes involve an integral over a solar model.
Solar Neutrinos 351
While helium and metal diffusion increases the neutrino fluxes, the
changes are roughly within the claimed errors of previous standard solar
model predictions. Many improvements of input physics over the years
have left the neutrino flux predictions of Bahcall and his collaborators
surprisingly stable over 25 years (Bahcall 1989, 1995). Bahcall (1994)
has compiled the central temperature predictions from a heterogeneous
set of 12 standard solar models without diffusion calculated by different
authors since 1988. The temperature predictions are almost uniformly
distributed on the interval 15.40−15.72 × 106 K, i.e. these authors agree
with each other on the value 15.56×106 K within ±1%.
Thus, in spite of differences in detail there exists a broad consensus
on what one means with a standard solar model. Therefore, the neu-
trino fluxes of the Bahcall and Pinsonneault (1995) model with element
diffusion (Tab. 10.1 and Fig. 10.1) can be taken to be representative.
The general agreement on a standard solar model does not guarantee, of
course, that there might not exist problems related to incorrect stan-
dard assumptions or incorrect input parameters common to all such
models.
correlation between Tc and the neutrino fluxes which allows one to un-
derstand the impact of certain modifications of a solar model on the
neutrino fluxes via their impact on Tc . Bahcall (1989) found
−1.2
T for pp,
c
Neutrino Flux ∝ Tc8 for 7 Be, (10.5)
Tc18 for 8 B.
b) Beryllium-Proton Reaction
A dominating uncertainty for the important flux of boron neutrinos
arises from the cross section 7 Be + p → 8 B + γ which plays no role
whatsoever for the energy generation in the Sun because the PPIII
termination of the pp chain is extremely rare. Therefore, a modification
of this cross section has no impact on the structure of the Sun and
thus no other observable consequence but to modify the high-energy
neutrino flux.
The cross section for this reaction is parametrized for low energies
in the usual form with an astrophysical S-factor
η = Z1 Z2 e2 v −1 . (10.7)
Here, Z1,2 e are the charges of the reaction partners, v their relative
velocity, and E their CM kinetic energy. The S-factor is expected to
be essentially constant at low energies unless there is a resonance near
threshold.
The six “classical” measurements of the 7 Be + p → 8 B + γ reaction
are referenced in Tab. 10.3. In the Sun, the most effective energy range
is around E = 20 keV, far in the tail of the thermal distributions of
the reaction partners, but still far below the lowest laboratory energies
of around 120 keV in the experiments of Kavanagh et al. (1969) and
Filippone et al. (1983). Therefore, one must extrapolate the factor
356 Chapter 10
10.3 Observations
10.3.1 Absorption Reactions for Radiochemical Experiments
The longest-running solar neutrino experiment is the Homestake chlo-
rine detector which is based on a reaction proposed by Pontecorvo
(1948) and Alvarez (1949),
νe + 37 Cl → 37 Ar + e (threshold 0.814 MeV). (10.9)
The two other data-producing radiochemical experiments use gallium
as a target according to the reaction
νe + 71 Ga → 71 Ge + e (threshold 0.233 MeV). (10.10)
Because of its low threshold, the gallium experiments can pick up the
solar pp neutrino flux. The absorption cross sections as a function of
358 Chapter 10
10−46 cm2 .
7 8 13 15
pp pep Be B N O
37
Cl 0 16 2.4 10,900 1.7 6.8
71
Ga 11.8 215 73.2 24,300 61.8 116
Table 10.5. Predicted absorption rate (in SNU) by 37 Cl for different solar
source reactions. BP92 give “theoretical 3σ” uncertainties.
7 8 13 15
Diffusion pp pep Be B N O Total
BP95 He, metals 0.0 0.2 1.2 7.2 0.1 0.4 9.3+1.2
−1.4
BP92 He 0.0 0.2 1.2 6.2 0.1 0.3 8.0 ± 3.0
BP92 — 0.0 0.2 1.2 5.5 0.1 0.2 7.2 ± 2.7
TL93 — 0.0 0.2 1.1 4.6 0.1 0.2 6.4 ± 1.4
The recognized systematic errors are 1.5% for the extraction ef-
ficiency, 3% for the proportional counter efficiency, 3% for the cos-
mic-ray background, 5% for the neutron background, and 2% for the
proportional counter background, which amounts to a total of 7%
or 0.18 SNU.
The measured counting rate for the individual runs at Homestake
is shown in Fig. 10.8. The current global best-fit average for the solar
neutrino flux measurement, derived from runs 18−124 is (Lande 1995)
Fig. 10.9. Solar neutrino flux at the gallium detectors (νe 71 Ga → 71 Ge e).
The shaded bands indicate a 1σ uncertainty of the global best-fit average
neutrino fluxes (statistical and systematic errors added in quadrature). In
the SAGE data, formally negative fluxes are forced to zero. In the right
panels, the bin size is 20 SNU.
364 Chapter 10
Table 10.6. Predicted absorption rate (in SNU) by 71 Ga for different solar
source reactions. BP92 give “theoretical 3σ” uncertainties.
7 8 13 15
Diffusion pp pep Be B N O Total
BP95 He, metals 69.7 3.0 37.7 16.1 3.8 6.3 137+8
−7
BP92 He 70.8 3.1 35.8 13.8 3.0 4.9 131.5+21
−17
BP92 — 71.3 3.1 32.9 12.3 2.7 4.3 127+19
−16
TL93 — 71.1 3.0 30.9 10.8 2.4 3.7 122.5 ± 7
Table 10.7. Coefficients in Eq. (10.15) for elastic neutrino electron scattering.
Flavor A B C
νe (CV + CA + 2)2 (CV − C A )2 (CV + 1)2 − (CA + 1)2
νe (CV − CA ) 2 (CV + CA + 2)2 (CV + 1)2 − (CA + 1)2
νµ,τ (CV + C A )2 (CV − C A )2 CV2 − CA2
ν µ,τ (CV − CA ) 2 (CV + CA ) 2 CV2 − CA2
CV = − 12 + sin2 ΘW ≈ −0.04, CA = − 12 .
366 Chapter 10
For Eν > 5 MeV, i.e. for energies above the detection threshold
of Kamiokande and Superkamiokande, the total cross section is well
approximated by
1 for νe ,
1 for ν e ,
G2F me Eν Eν 2.4
σνe ≈ (A + 13 B) = 9.5×10−44 cm2 × 1
2π 10 MeV
6.2 for νµ,τ ,
1
7.1
for ν µ,τ .
(10.17)
device and thus a true “neutrino telescope.” Thus, for the first time
the Kamiokande detector actually proved that neutrinos are coming
from the direction of the Sun. The first main publication of these
results was by Hirata et al. (1991) from a data sample of 1040 live de-
tector days, taken between 1987 and 1990; an update including data
until July 1993 (a total of 1670 live detector days) was given by Suzuki
(1995). The angular distribution of the registered electrons relative to
the direction of the Sun is shown in Fig. 10.11. Even though there
remains an isotropic background, probably from radioactive impurities
in the water, the solar neutrino signal beautifully shows up in these
measurements.
In elastic neutrino-electron scattering, the energy distribution of the
kicked electrons is nearly flat so that the spectral shape of the incident
neutrino flux is only indirectly represented by the measured electron
spectrum. In Fig. 10.12 the electron recoil spectrum is shown for an
incident spectrum of 8 B neutrinos. A water Cherenkov detector can
resolve the energy of the charged particles from the intensity of the
measured light which for low-energy electrons is roughly proportional
to the energy. Therefore, the electron recoil spectrum from the interac-
tion with solar neutrinos can be resolved at Kamiokande. The measured
shape relative to the theoretically expected one is shown in Fig. 10.13,
arbitrarily normalized at Eν = 9.5 MeV (Suzuki 1995). Within statis-
tical fluctuations the agreement is perfect.
368 Chapter 10
Even though the Sun is thought to produce only νe ’s one may spec-
ulate that some of them are converted into ν e ’s on their way to Earth
by spin-flavor oscillations in magnetic fields (Sect. 8.4) or by matter-
induced majoron decays (Sect. 6.8). The ν e ’s would cause a signal by
the reaction ν e p → ne+ with an isotropic angular distribution. The
remaining measured isotropic background shown in Fig. 10.11 can thus
Solar Neutrinos 369
Fig. 10.13. Measured over expected spectrum of recoil electrons from solar
neutrinos at Kamiokande, arbitrarily normalized to unity at Eν = 9.5 MeV
(Suzuki 1995).
Fig. 10.14. Limit on a solar ν e flux relative to 8 B solar νe ’s from the isotropic
background shown in Fig. 10.11. The data used for this result extend until
August 1992 (Suzuki 1993).
Table 10.8. Predicted flux of 8 B neutrinos at Earth. For BP92, the uncer-
tainty is a “theoretical 3σ error.”
Model Diffusion 8
B Flux [106 cm−2 s−1 ]
BP95 He, metals 6.6+0.9
−1.1
BP92 He 5.7 ± 2.4
BP92 — 5.1 ± 2.2
TL93 — 4.4 ± 1.1
10.3.5 Summary
The main features and results of the solar neutrino experiments dis-
cussed in this section, and of near-future experiments to be discussed
in Sect. 10.9, are summarized in Tab. 10.9. More technical aspects can
be found in the original papers quoted in this section and in previous
papers by the referenced authors. Overviews of many experimental as-
pects can be found in Bahcall (1989), Davis, Mann, and Wolfenstein
(1989), and Koshiba (1992). The two theoretical predictions shown
are somewhat extreme in that TL93 yields lowish neutrino fluxes when
compared with BP92 (no diffusion), while BP95 with the inclusion of
helium and metal diffusion is presently at the upper end of what is
being predicted on the basis of standard solar models. It is clear, of
course, that including diffusion would also increase the TL93 fluxes.
55
The refractive term in the MSW conversion probability from convective currents
could play a role and could be coupled to solar activity (Haxton and Zhang 1991).
However, extreme conditions are required to explain the observed effect.
Solar Neutrinos 375
and for the central 14◦ × 14◦ which is the region most significant for the
neutrino flight path to us. The disk-centered magnetic-field cycle lags
full-disk indicators by about 1 y. Oakley et al. (1994) found a signifi-
cance level of 0.001% for an anticorrelation between the Homestake rate
and the disk-centered magnetic flux. Including the flux from increas-
ingly higher latitudes reduced the significance level of the correlation.
With only high-latitude magnetic information the correlation was lost.
Another test of time variation is a comparison as in Fig. 10.7 be-
tween the expected and measured distribution of argon production
rates. A time variation would broaden the measured distribution (black
histogram) relative to the expectation for a constant neutrino flux
(shaded histogram). A certain degree of broadening is certainly com-
patible with Fig. 10.7, but a precise statistical analysis does not seem
to be available at the present time. At any rate, the rank-ordering re-
sult does not specify the amplitude of a time varying signal relative to a
constant base rate while the width of the distribution in Fig. 10.7 would
be sensitive mostly to this amplitude. Therefore, the two methods yield
rather different information concerning a possible time variation.
The gallium experiments have not been running long enough to say
much about a time variation on the time scale of several years. Also,
there does not seem to be a significant time variation in the Kamiokande
data between 1987 and 1992 which covers a large fraction of the cur-
rent solar cycle No. 22. Notably, there is no apparent (anti)correlation
with sunspot number (Suzuki 1993). However, a correlation with disk-
centered magnetic indicators may yield a different result—according to
the analysis of Oakley et al. (1994) the use of a disk-centered indicator
with its inherent time-lag relative to full-disk spot counts is crucial for
the strong anticorrelation at Homestake.
The constancy of the Kamiokande rate, if confirmed over a more
extended period, severely limits a conjecture that the nuclear energy
generating region in the Sun was not constant (e.g. Raychaudhuri 1971,
1986). Because the Kamiokande detector is mostly sensitive to the
boron neutrinos with their extreme sensitivity to temperature, they
would be expected to show the most extreme time variations if the
solar cycle was caused by processes in the deep interior rather than by
dynamo action in the convective surface layers.
10.4.4 Summary
There is no indication for a day-night variation of the solar neutrino
flux at Kamiokande and none for a seasonal variation at Kamiokande,
Solar Neutrinos 377
Even if one ignores for the moment a possible time variation of the
Homestake neutrino measurements there remain several “solar neutrino
problems.” All of the solar neutrino flux measurements discussed in
Sect. 10.3 exhibit a deficit relative to theoretical predictions. The sim-
plest case to interpret is that of Kamiokande because it is sensitive only
to the boron flux. In this case the most uncertain input parameter is
the cross section for the reaction p7 Be → 8 Bγ which enters the flux pre-
diction as a multiplicative factor, independently of other details of solar
modelling. As discussed in Sect. 10.2.3, the astrophysical S-factor for
this reaction depends on theoretical extrapolations to low energies of
experimental data which themselves seem to exhibit relatively large sys-
tematic uncertainties. Therefore, one may turn the argument around
and consider the solar neutrino flux measurement at Kamiokande as
another determination of S17 (0).
According to Eq. (10.20) the measured flux is 2.89 × (1 ± 0.14) in
units of 106 cm−2 s−1 while the prediction is 4.4×(1±0.21)×S17 (0)/22.4
for TL93 (Turck-Chièze and Lopes 1993) where S17 is understood in
units of eV b. For the BP95 (Bahcall and Pinsonneault 1995) it is
6.6 × (1 ± 0.15) × S17 (0)/22.4 where I have used an average symmetric
378 Chapter 10
Table 10.10. Measured vs. predicted 7 Be and CNO neutrino fluxes (in SNU).
Homestake GALLEX/
(37 Cl) SAGE (71 Ga)
Measurements:
Total 2.55 ± 0.25 77 ± 10
8
B (inferred from 3.15 ± 0.44 7±1
Kamiokande)
pp + pep 0.20 ± 0.01 74 ± 1
(calculated)
Remainder −0.8 ± 0.5 −4 ± 10
BP95 prediction:a
7
Be 1.24 37.7
CNO 0.48 10.1
Total 1.72 47.8
deficits in all detectors may give us the first indication for neutrino
oscillations and thus for nonvanishing neutrino masses and mixings.
Because of the statistically high significance of the anticorrelations
with solar activity of the Homestake results it is not entirely obvious
how to proceed with a quantitative test of the oscillation hypothesis
which can cause only a day-night or semiannual time variation. In the
present section a possible long-term variability of the solar neutrino
flux or its detection methods is ignored, i.e. the apparent anticorrela-
tion with solar activity at Homestake is considered to be a statistical
fluctuation. In Sect. 10.7 a possible explanation in terms of neutrino
interactions with the solar magnetic field is considered.
One may take the opposite point of view that the variability at
Homestake is a real effect, and that it is not related to neutrino mag-
netic moments because that hypothesis requires fairly extreme values
for the dipole moments and solar magnetic fields (Sect. 10.7). An in-
terpretation of the data in terms of neutrino oscillations then becomes
difficult because one admits from the start that unknown physical ef-
fects are either operating in the Sun, in the intervening space, or in
the detectors. One could argue perhaps that the apparent variability
of the Homestake data in itself was evidence that something was wrong
with this experiment. In this case one could still test the hypothesis
of neutrino oscillations under the assumption that the signal recorded
at Homestake was spurious in which case one must discard the entire
data set, not only parts of it as has sometimes been done. Ignoring
the Homestake data does not solve the solar neutrino problem, but its
significance is reduced.
Still, as no one has put forth a plausible hypothesis for a specific
problem with the Homestake experiment it is arbitrary to discard the
data. Admittedly, it is also arbitrary to consider the time variation
spurious even though standard statistical methods seem to reveal a
highly significant anticorrelation with solar activity.
sin2 2θ = 0.6,
∆m2ν = 2×10−5 eV2 , (10.24)
sin2 2θ = 0.6×10−2 ,
∆m2ν = 0.6×10−5 eV2 . (10.25)
Fig. 10.19. Allowed range of neutrino masses and mixing angles in the neu-
trino experiments if the flux deficit relative to the Bahcall and Pinsonneault
(1995) solar model is interpreted in terms of neutrino oscillations. The ex-
perimental data include all summarized in Sect. 10.3. (Plot adapted from
Hata and Haxton 1995.)
58
Recent detailed investigations were performed by Akhmedov, Lanza, and Petcov
(1993, 1995), Krastev (1993), Nunokawa and Minakata (1993), Guzzo and Pulido
(1993), and Pulido (1993, 1994). For a review of earlier works see Pulido 1992.
Solar Neutrinos 389
If some form of neutrino oscillations are the explanation for the mea-
sured solar flux deficits relative to standard predictions, how are we
ever going to know for sure? One needs to measure a signature which
is characteristic only for neutrino oscillations. The most convincing
case would be a measurement of the “wrong-flavored” neutrinos, i.e.
the νµ or ντ appearance rather than the νe disappearance. Other clear
signatures would be a deformation of the 8 B spectrum or a diurnal or
seasonal flux variation.
The latter cases can be very well investigated with the Superkamio-
kande detector which is scheduled to begin its operation in April of
1996. It is a water Cherenkov detector like Kamiokande, with about
20 times the fiducial volume. With about twice the relative coverage
of the surface area with photocathodes and a detection threshold as
low as 5 MeV it will count about 30 events/day from the solar boron
neutrino flux, as opposed to about 0.3 events/day at Kamiokande.
Fig. 10.20. Expected signal at night relative to the average daytime sig-
nal in a water Cherenkov detector as a function of the angle between the
Sun and the detector nadir. The large-angle example is for sin2 2θ = 0.7
and ∆m2ν = 1×10−5 eV2 , the small-angle case for sin2 2θ = 0.01 and
∆m2ν = 0.3×10−5 eV2 . Also shown are the existing Kamiokande measure-
ments and the expected Superkamiokande error bars after 1 month and 1 year
of running, respectively. (Adapted from Suzuki 1995.)
Solar Neutrinos 391
Fig. 10.21. Spectral distortion of the recoil electrons from the primary boron
neutrinos in a water Cherenkov detector. The ratio relative to the standard
spectrum is arbitrarily normalized at an electron kinetic energy of Te =
10 MeV. For the large-angle example (solid line) the assumed mass-square
difference is ∆m2ν = 2×10−5 eV2 , for the small-angle examples (broken lines)
it is ∆m2ν = 0.6×10−5 eV2 . The anticipated error bars after 5 years of
running Superkamiokande are also indicated for two energies. (Adapted
from Krastev and Smirnov 1994.)
392 Chapter 10
10.9.3 BOREXINO
Superkamiokande and SNO are both limited to a measurement of the
boron neutrino flux because of their relatively high detection thresholds.
If the boron flux is partly or mostly suppressed by a low S17 factor or
a low central solar temperature instead of neutrino oscillations these
experiments may have difficulties at identifying oscillations which would
still be indicated by the missing beryllium neutrino flux. Therefore, it is
interesting that another experiment (BOREXINO60 ) is being prepared
which would be sensitive dominantly to the beryllium neutrinos.
The main detection reaction is elastic ν-e scattering as in the light-
water Cherenkov detectors. However, the kicked electron is detected by
virtue of scintillation light rather than Cherenkov radiation, the former
60
The name of this experiment is derived from BOREX (boron solar neutrino
experiment), a proposed detector that was to use 11 B as a target (Raghavan, Pak-
vasa, and Brown 1986; see also Bahcall 1989). For a practical implementation it
was envisaged to use a boron loaded liquid scintillator (e.g. Raghavan 1990); be-
cause of the relatively small size of this detector the Italian diminutive BOREXINO
(baby BOREX) emerged. Ultimately, the idea of using a borated scintillator was
dropped entirely, leaving boron only in the name of the experiment. Confusingly,
then, BOREXINO is unrelated to a boron target, and also unrelated to the solar
boron neutrinos because the experiment is designed to hunt the beryllium ones.
394 Chapter 10
yielding about 50 times more light at the relevant energies below about
1 MeV. Naturally, as a target one needs to use an appropriate scintilla-
tor rather than water. The advantage of a lowered threshold is bought
at the price of losing all directional information. However, because of
the good energy resolution the monochromatic beryllium neutrinos at
862 keV should be clearly detectable as a distinct shoulder in the energy
spectrum of the recoil electrons. Optimistically, a scintillation detector
could have a threshold as low as Eν = 250 keV.
The main challenge at implementing this method is to lower the
radioactive contamination of the scintillator, its vessel, and the sur-
rounding water bath to an unprecedented degree of purity. For exam-
ple, the allowed mass fraction of 238 U of the scintillator is less than
about 10−16 g/g. The feasibility of this method is currently being stud-
ied at the CTF (Counting Test Facility) experiment, located in the
Gran Sasso underground laboratory. Assuming a positive outcome,
BOREXINO would be built, consisting of 300 tons of scintillator, sur-
rounded by 3000 tons of water. Optimistically, data taking with this
facility could commence in 1997.
10.9.5 Summary
The hypothesis of neutrino flavor oscillations is strongly supported by
the results of all existing solar neutrino experiments. With the new
generation of detectors which will begin to take up operation in 1996
it looks plausible that nonstandard neutrino properties can be firmly
established on the basis of the solar neutrino flux before the mille-
nium ends.
Chapter 11
Supernova Neutrinos
395
396 Chapter 11
Fig. 11.1. Schematic picture of the core collapse of a massive star (M > ∼
8 M! ), of the formation of a neutron-star remnant, and the beginning of a
SN explosion. There are four main phases numbered 1−4 above the plot:
1. Collapse. 2. Prompt-shock propagation and break-out, release of prompt
νe burst. 3. Matter accretion and mantle cooling. 4. Kelvin-Helmholtz
cooling of “protoneutron star.” The curves mark the time evolution of several
characteristic radii: The stellar iron core (RFe ). The “neutrino sphere” (Rν )
with diffusive transport inside, free streaming outside. The “inner core”
(Ric ) which for t <∼ 0.1 s is the region of subsonic collapse, later it is the
settled, compact inner region of the nascent neutron star. The SN shock
wave (Rshock ) is formed at core bounce, stagnates for several 100 ms, and
is revived by neutrino heating—it then propagates outward and ejects the
stellar mantle. The shaded area is where most of the neutrino emission
comes from; between this area and Rν neutrinos still diffuse, but are no
longer efficiently produced. (Adapted from Janka 1993.)
Neutrino trapping has the effect that the lepton number fraction
YL is nearly conserved at the value Ye which obtains at the time of
trapping. However, electrons and electron neutrinos still interconvert
(β equilibrium), causing a degenerate νe sea to build up. The core of
a collapsing star is the only known astrophysical site apart from the
early universe where neutrinos are in thermal equilibrium. It is the
only site where neutrinos occur in a degenerate Fermi sea as the early
universe is thought to be essentially CP symmetric with equal numbers
of neutrinos and antineutrinos to within one part in 109 . When neutrino
trapping becomes effective, the lepton fraction per baryon is YL ≈ 0.35,
398 Chapter 11
not much lower than the initial iron-core value, i.e. not much of the
lepton number is lost during infall. The conditions of chemical and
thermal equilibrium dictate that neutrinos take up only a relatively
small part (about 14 ) of the lepton number (Appendix D.2). A typical
YL profile after collapse is shown in Fig. 11.2
The collapse is intercepted when the inner core reaches nuclear den-
sity (ρ0 ≈ 3×1014 g cm−3 ), a point where the equation of state stiffens.
Because the inner core collapse is subsonic, the information about the
central condition spreads throughout, i.e. the collapse of the entire ho-
mologous core slows down. However, this information cannot propagate
beyond the sonic point at the edge of the inner core which now encom-
passes about 0.8 M! . As material continues to fall onto the inner core
at supersonic velocities a shock wave builds up at the sonic point which
is at the edge of the inner core, not at its center. As more material
moves in, more and more energy is stored in this shock wave which
almost immediately begins to propagate outward into the collapsing
outer part of the iron core—see the thick solid line in Fig. 11.1. As-
suming that enough energy is stored in the shock wave it will eventually
eject the stellar mantle outside of what was the iron core. The rebound
or “bounce” of the collapse turns the implosion of the core into an
explosion of the outer star—a SN occurs.
This “bounce and shock” scenario of SN explosions was first pro-
posed by Colgate and Johnson (1960) and then elaborated by a number
of authors (see Brown, Bethe, and Baym 1982 and references therein).
In practice, however, the story of SN explosions appears to be more
complicated than this “prompt explosion scenario.” Neutrino losses
and the dissociation of the iron material through which the shock wave
propagates dissipate much of the shock’s energy so that in typical calcu-
lations it stalls and eventually recollapses. It is currently believed that
the energy deposition by neutrinos revives the shock wave, leading to
the “delayed explosion scenario” detailed in Sect. 11.1.3 below.
Fig. 11.3. Schematic neutrino “lightcurves” during the phases of (1) core
collapse, (2) shock propagation and shock breakout, (3) mantle cooling and
accretion, and (4) Kelvin-Helmholtz cooling. (Adapted from Janka 1993.)
its lepton number during the νe burst at shock break-out. This outer
part settles within the first 0.5−1 s after core bounce, emitting most
of its energy in the form of neutrinos. Also, more material is accreted
while the shock wave stalls. As much as a quarter of the expected total
amount of energy in neutrinos is liberated during this phase (No. 3 in
Figs. 11.1 and 11.3).
Meanwhile, the stalled shock wave has managed to resume its out-
ward motion and has begun to eject the overburden of matter. There-
fore, the protoneutron star after about 0.5−1 s can be viewed as a star
unto itself with a radius of around 30 km which slowly contracts and
cools by the emission of (anti)neutrinos of all flavors, and at the same
time deleptonizes by the loss of νe ’s. After 5−10 s it has lost most
of its lepton number, and slightly later most of its energy. This is
the “Kelvin-Helmholtz cooling phase,” marked as No. 4 in Figs. 11.1
and 11.3. Afterward, the star has become a proper neutron star whose
small lepton fraction is determined by the condition of a vanishing
neutrino chemical potential (Appendix D.2), and whose further cooling
history has been discussed in Sect. 2.3.
Immediately after collapse the protoneutron star is relatively cold
(see the t = 0 curve in Fig. 11.2). Half or more of the energy to be ra-
diated later is actually stored in the degenerate electron Fermi sea with
typical Fermi momenta of order 300 MeV. The corresponding degener-
Supernova Neutrinos 401
ate neutrino sea has its Fermi surface at around 200 MeV. The lepton
number profile (lower panel of Fig. 11.2) has a step-like form, with the
step moving inward very quickly when the shock breaks through the
neutrino sphere. The quick recession of the lepton profile during the
first 0.5 s represents deleptonization during the mantle cooling phase.
After this initial phase, however, the bloated outer part of the star
has settled; it is more difficult for neutrinos to escape from this compact
object. Still, the steep gradient of lepton number drives an outward dif-
fusion of neutrinos which move toward regions of lower Fermi momen-
tum and thus, of lower degeneracy energy. Therefore, they “downscat-
ter,” releasing most of the previous electron and neutrino degeneracy
energy as heat. Hence near the edge of the lepton number step the
medium is heated efficiently. In Fig. 11.2 it is plainly visible that the
temperature maximum of the medium is always in the region of the
steepest lepton number gradient.62 Therefore, the medium first heats
near the core surface, and then the temperature maximum moves in-
ward until it has reached the center. At this time the core is entirely
deleptonized; it continues to cool, the temperature maximum at the
center drops to obscurity.
The neutrino radiation leaving the star has typical energies in the
10 MeV range, compared with a 200 MeV neutrino Fermi energy in the
interior. Therefore, the loss of lepton number by itself is associated
with relatively little energy. Put another way, only a small excess of νe
over ν e is needed to carry away the lepton number. The total energy
is carried away in almost equal parts by each (anti)neutrino flavor.
Fig. 11.4. Unsuccessful (upper panel) and successful (lower panel) SN explo-
sion. In each case, the location of several mass shells is shown as a function
of time. The thick shaded line indicates the location of the shock. The only
difference between the two cases is the adjusted neutrino luminosity from a
central source; it was chosen as Lν = 2.10×1052 erg s−1 (upper panel) and
Lν = 2.20×1052 erg s−1 (lower panel), respectively. In the upper case, which
is just below threshold for a successful explosion, the shock displays an in-
teresting oscillatory behavior. (Curves courtesy of H.-T. Janka, taken from
Janka and Müller 1993a.)
404 Chapter 11
Fig. 11.5. Entropy contours between neutron star and shock wave in a 2-
dimensional calculation of a SN explosion. The entropy per nucleon is shown
in contours at equal steps of 0.5 kB between 5 and 16 kB , and in steps of 1 kB
between 16 and 23 kB . This snapshot represents model T2c of Janka and
Müller (1995b) at t = 377 ms after bounce. (Original of the figure courtesy
of H.-T. Janka.)
Supernova Neutrinos 405
Hayes, and Fryxell 1995). For the first time 2- and 3-dimensional cal-
culations have become possible. They reveal a large-scale convective
overturn (Fig. 11.5) which helps at revitalizing the shock because it
brings hot material from depths near the neutrino sphere quickly up
to the region immediately behind the shock, and cooler material down
to the neutrino sphere where it absorbs energy from the neutrino flow.
Successful explosions can be obtained for amounts of neutrino heating
where 1-dimensional calculations did not succeed. The sharp transition
between failed and successful explosions as a function of neutrino heat-
ing that was found in 1-dimensional calculations (Fig. 11.4) is smoothed
out, but neutrino heating still plays a pivotal role at obtaining a suc-
cessful and sufficiently energetic explosion.
In summary, then, the current standard picture of SN explosions is
a modification and synthesis of the Colgate and Johnson (1960) shock-
driven and the Colgate and White (1966) neutrino-driven explosions.
At the present time there may still remain a quantitative problem at
obtaining enough neutrino energy deposition behind the shock wave to
guarantee a successful and sufficiently energetic explosion. It remains
to be seen if this scenario withstands the test of time, or if a novel
ingredient will have to be invoked in the future.
11.1.4 Nucleosynthesis
The universe began in a hot “big bang” which allowed for the formation
of nuclei from the protons and neutrons originally present in thermal
equilibrium with the ambient heat bath. The primordial abundances
“froze out” at about 22−24% helium, the rest hydrogen, and a small
trace of other light elements such as lithium. The present-day distri-
bution of elements was bred from this primeval mix mostly by nuclear
processes in stars; they eject some of their mass at the end of their lives
(Chapter 2), returning processed material to the interstellar medium
from which new stars and planets are born.
However, the normal stellar burning processes can produce elements
only up to the iron group which have the largest binding energy per
nucleon. Thus, the heavy elements must have been produced by dif-
ferent processes at different sites. It has long been thought that nu-
clei with A >∼ 70 were predominantly made by neutron capture, no-
tably the s- and r- (slow and rapid) processes (Burbidge et al. 1957;
Cameron 1957; Clayton 1968; Meyer 1994). The site for the occur-
rence of the r-process has remained elusive for the past three decades,
although many different suggestions have been made. The crux is that
406 Chapter 11
i.e. typically 'Eνe ( ≈ 23 'Eν e ( and 'Eν ( ≈ 53 'Eν e ( for the other flavors.
The number fluxes of the nonelectron flavors are smaller than those
of ν e because the energy is found to be approximately equipartitioned
between the flavors: the total Eνe +ν e lies between 13 and 12 of Eb . Simi-
larly, the number flux of νe is larger than that of ν e (the lepton number
is carried away in νe ’s!) so that, again, the energy is approximately
equipartitioned between νe and ν e . The total Eν e is found to lie between
1
6
and 14 of Eb . The SN 1987A observations were almost exclusively sen-
sitive to the ν e flux. The total Eν e inferred from these measurements
65
See, e.g., Burrows and Lattimer (1986); Bruenn (1987); Mayle, Wilson, and
Schramm (1987); Burrows (1988); Janka and Hillebrandt (1989a,b); Myra and
Bludman (1989); Myra and Burrows (1990). For reviews see Cooperstein (1988)
and Burrows (1990a,b).
Supernova Neutrinos 409
bounce accretion, and the temperature profile of the core after col-
lapse.
As expected, a soft EOS leads to a large amount of binding energy
and thus to large integrated neutrino luminosities; a large core mass
or large postbounce accretion rate has a similar effect. A soft EOS
leads to relatively high temperatures during deleptonization, causing
large neutrino opacities and thus long emission time scales. If the EOS
is too soft, or the core mass too large, the final configuration is not
stable and collapses, presumably to a black hole. This must not occur
too early to avoid conflict with the duration of the observed SN 1987A
Fig. 11.7. Luminosity and temperature of the ν e flux from the protoneutron
star model 55 of Burrows (1988) which is based on a “stiff equation of state,”
an initial baryonic core mass of 1.3 M! , and an accretion of 0.2 M! within
the first 0.5 s. The dotted line in the upper panel indicates a t−1 behavior,
in the lower panel it indicates e−t/4τ with τ ≈ 10 s. The neutrino spectral
distribution was taken to be thermal.
Supernova Neutrinos 413
signal (Sect. 11.3). However, in some cases studied by Keil and Janka
(1995) with an EOS including hyperons this final collapse occurs so
late (at 8 s after bounce in one example) that black-hole formation is
difficult to exclude on the basis of the SN 1987A observations, notably
as no pulsar has yet been found there.
Needless to say, with so many parameters to play it is not difficult
to find combinations of EOS, core mass, accretion rate, and initial
temperatures which fit the observed SN 1987A signal well within the
statistical uncertainties of the observations. An example is model 55
of Burrows (1988) which is based on a “stiff EOS,” an initial baryonic
core mass of 1.3 M! , and an accretion of 0.2 M! within the first 0.5 s.
The evolution of the effective ν e luminosity and temperature is shown
in Fig. 11.7. After about 1 s the decay of the temperature is fit well by
an exponential e−t/4τ with τ ≈ 10 s while the decay of the luminosity
is poorly fit by an exponential; it decays approximately as t−1 after 1 s.
For later reference, the time-integrated flux (fluence) of this model is
shown in Fig. 11.8.
It must be stressed that the “cooling behavior” (decrease of the av-
erage ν e energy) shown in Fig. 11.7 may not be generic at early times.
Initially the star is quite bloated, and relatively cold. Therefore, the
time) on 23 February 1987 while the first evidence for optical brighten-
ing was found at 10:38 UT on plates taken by McNaught (1987)—see
Fig. 11.9.
The main neutrino observations come from the Irvine-Michigan-
Brookhaven (IMB) and the Kamiokande II water Cherenkov detectors,
facilities originally built to search for proton decay, while a less sig-
nificant measurement is from the Baksan Scintillator Telescope (BST).
A likely spurious observation is from the Mont Blanc Liquid Scintilla-
tor Detector (LSD). It preceded the other observations by about 5 h,
with no contemporaneous signal at Mont Blanc with the other signals,
and no contemporaneous signal at the other detectors with the Mont
Blanc event. The Mont Blanc detector was built to search for neutrinos
from core collapse supernovae, except that it was optimized for galac-
tic events within a distance of about 10 kpc. The neutrino output of
a normal SN in the LMC could not have caused an observable signal
at Mont Blanc; the reported events probably represent a background
fluctuation.
Koshiba (1992) has given a lively account of the exciting and ini-
tially somewhat confusing story of the neutrino measurements and their
interpretation. Early summaries of the implications for astrophysics
and particle physics of the neutrino and electromagnetic observations
were written, for example, by Schramm (1987), Arnett et al. (1989),
and Schramm and Truran (1990). A more recent review of SN 1987A
is McCray (1993). A nontechnical overview was provided in a book by
Murdin (1990).
For the present purposes the bottom line is that SN 1987A broadly
confirmed our understanding of SN physics as outlined in Sect. 11.1.
A remaining sore point is the lack of a pulsar observation in the SN
remnant so that one may continue to speculate that a black hole has
formed in the collapse.
where GF is the Fermi constant and cos2 θC ≈ 0.95 refers to the Cabibbo
angle. The charged-current vector and axial-vector weak coupling con-
stants are CV = 1 and CA = 1.26, and δ incorporates small corrections
from recoil, Coulomb, radiative and weak magnetism corrections (Vo-
gel 1984). Further, pe and Ee refers to the positron momentum and
energy. Ignoring recoil effects, the latter is Ee = Eν − mn + mp ≈
Eν − 1.3 MeV; the threshold is 1.8 MeV because the minimum Ee is me .
A general expression for the νe 16 O cross section is much more com-
plicated. A simple approximation, taking only the 2− state of the 16 F
nucleus, is
Fig. 11.10. Total cross sections for the measurement of neutrinos in a water
Cherenkov detector according to Eqs. (10.17), (11.6), and (11.7). The curves
refer to the total cross section per water molecule so that a factor of 2 for
protons and 10 for electrons is already included.
Supernova Neutrinos 417
All of the relevant cross sections per water molecule are shown in
Fig. 11.10 as a function of Eν . The curves incorporate a factor of 2 for
proton targets (two per H2 O), and a factor of 10 for electrons (ten per
H2 O). Above its threshold, the νe 16 O cross section rises very fast; it is
then the dominant detection process for νe ’s. Still, the ν e p reaction is
the absolutely dominant mode of observing SN neutrinos.
The BST detector is filled with an organic scintillator based on
“white spirit” Cn H2n+2 with n ≈ 9. The dominant detection reaction is
also ν e p → ne+ . In addition, elastic scattering on electrons is possible,
and the process νe 12 C → 12 N e− occurs for Eνe > ∼ 30 MeV.
The trigger efficiencies relevant for the three detectors are shown
as a function of the e± energy in Fig. 11.11. Analytic fit formulae to
these curves were given by Burrows (1988) for IMB and Kamiokande.
IMB reports a dead time of 13% during the SN burst (Bratton et al.
1988); the IMB curve includes a factor 0.87 to account for this effect.
The fiducial volume of Kamiokande II relevant for the SN 1987A obser-
vations was 2,140 tons, for IMB 6,800 tons, and for BST 200 tons. It
corresponds to a target of 1.43×1032 protons at Kamiokande, 4.6×1032
at IMB, and 1.88×1031 at BST.
With the ν e p cross section of Eq. (11.6) and the efficiency curves
of Fig. 11.11 one may compute a prediction for the number of events
Table 11.1. Neutrino burst at the Kamiokande detector (Hirata et al. 1988).
The time is relative to the first event at 7:35:35 ± 0:01:00 UT, 23 Feb. 1987.
The energy refers to the detected e± , not to the primary neutrino.
Table 11.2. Neutrino burst at the IMB detector (Bratton et al. 1988).
The time is relative to the first event at 7:35:41.374 ± 0:00:00.050 UT, 23
Feb. 1987. The energy refers to the detected e± , not to the primary neutrino.
Table 11.3. Neutrino burst at the Baksan detector (Alexeyev et al. 1987,
1988). The time is relative to the first event at 7:36:06.571+02.000
−54.000 UT, 23
Feb. 1987. The energy refers to the detected e± , not to the primary neutrino.
the final-state electron energy distribution is broad so that one can infer
only a lower limit to the ν energy.
The measured events at the Kamiokande (Hirata et al. 1987, 1988),
IMB (Bionta et al. 1987; Bratton et al. 1988), and BST (Alexeyev et al.
1987, 1988) detectors are summarized in Tabs. 11.1, 11.2, and 11.3.
The absolute timing at IMB is accurate to within ±50 ms while at
Kamiokande only to within ±1 min. At BST, the clock exhibited an
erratic behavior which led to an uncertainty of +2/−54 s. Within the
timing uncertainties the three bursts are contemporaneous and may
Fig. 11.14. SN 1987A neutrinos at Kamiokande and IMB, excluding the ones
which likely are due to background.
Supernova Neutrinos 423
so that τ is the decay time scale of the luminosity which varies with
the fourth power of the temperature according to the Stefan-Boltzmann
law. It should be noted, however, that numerical cooling calculations
do not yield exponential lightcurves. For example, the model shown in
Fig. 11.7 displays an exponential decline of the effective temperature,
but a power-law decline of the neutrino luminosity. Other calculations
even yield early heating and a constant temperature for some time (see
Burrows 1990b for an overview).
Of course, for the time-integrated spectrum the exponential cooling
law is just another assumption concerning the overall spectral shape.
For example, one easily finds that the average ν e energy of the time-
integrated spectrum is 'Eν e ( = 2.36 T0 if Fermi-Dirac distributions with
424 Chapter 11
mν e <
∼ 20 eV (11.10)
high-energy forward component; they are not fit well by the assumed
dominant detection process ν e p → ne+ which is supposed to yield an
isotropic positron distribution with no directional correlation with en-
ergy. Elastic νe scattering, however, does not fit the data well either
because it is too forward peaked, and anyhow it is disfavored by a small
cross section unless the flux of νµ,τ or ν µ,τ was extremely high.
However, because no plausible and/or viable nonstandard cause for
the observed events has been proposed one has settled for the interpre-
tation of a statistical fluctuation for the apparent anomalies. After all,
it is difficult to imagine a small sample drawn from any distribution
without some “anomalies” which are easy to overinterpret. Still, if a
reasonable alternative to the standard interpretation of the signal were
to come forth this topic would have to be reconsidered.
11.4.1 Overview
ber of νe ’s in the burst is of order 1056 , its duration of order 50 ms. Thus,
while it passes it represents a νe density of about 1032 cm−3 r7−2 which
exceeds the local electron density for r > 9
∼ 10 cm. However, the phase-
space distribution of the neutrinos is locally far from isotropic and so
the energy shift involves a factor '1 − cos Θ( where Θ is the angle be-
tween the “test neutrino” and a “background neutrino;” the average is
to be taken over all background neutrinos (Sect. 9.3.2). A typical angle
between two neutrinos moving within the burst at the same location
is given by the angle subtended by the neutrino sphere as viewed from
the relevant radial position, i.e. Θ ≈ R/r with R the radius of the neu-
trino sphere. For a large r one thus finds '1 − cos Θ( ≈ Θ2 ≈ (R/r)2 .
Then, with R ≈ 107 cm the effective neutrino density is approximately
1032 cm−3 r7−4 , a value which is always smaller than the electron density
Eq. (11.11). Therefore, in the present context one may ignore neutrino-
neutrino interactions. This will not be the case for the issue of r-process
nucleosynthesis (Sect. 11.4.5).
One may proceed to determine the MSW triangle as in Fig. 8.9
for solar neutrinos, except that there an exponential electron density
profile was used while now Eq. (11.11) pertains. Nötzold (1987) found
a conversion probability in excess of 50% if
Fig. 11.17. MSW triangle for the prompt νe burst from a stellar collapse. In
the shaded area the conversion probability exceeds 50% for Eν = 20 MeV,
assuming the electron density profile of Eq. (11.11). The MSW solutions
to the solar neutrino problem and the Kamiokande allowed range for solar
neutrinos are indicated (see Fig. 10.19).
11.4.3 Cooling-Phase ν e ’s
Fig. 11.18. Contours for the “swap fraction” p between the ν e and the ν µ
or ν τ fluxes from the protoneutron star cooling phase. The matter effect of
the stellar envelope and of the Earth are included. In (A) the Earth effect is
important; in this area p is an average over neutrino energies while otherwise
it does not depend on the energy. In (B) the exact contours depend on the
detailed matter distribution of the stellar envelope. Black areas indicate
the approximate mixing parameters which would explain the solar neutrino
problem. (Adapted from Smirnov, Spergel, and Bahcall 1994.)
436 Chapter 11
two detectors. In this case the contours refer to an average value 'p(.
In the shaded area (B) the contours are not independent of the detailed
matter distribution in the stellar envelope.
In Fig. 11.18 the large-angle MSW and the vacuum oscillation solu-
tions of the solar neutrino problem are indicated. If either one of them
is correct the measurable ν e spectrum in a detector is a substantial mix-
ture of different-flavor source spectra. Smirnov, Spergel, and Bahcall
(1994) argued on the basis of a joint analysis between the SN 1987A
signals at the Kamiokande and IMB detectors that p < 0.17−0.27 at
the 95% CL, depending on the assumed primary neutrino spectra. If
p were any larger, the expected spectra would be much harder than
has been observed. This analysis excludes the vacuum oscillation solu-
tion to the solar neutrino problem. Kernan and Krauss (1995) arrive
at the opposite conclusion that all mixing angles are permitted by the
SN 1987A signal, and that sin2 2θ = 0.45 is actually a favored value.
Fig. 11.19. Typical density profile for a SN model at 0.15√ s after the core
bounce (Fuller et al. 1992). The right-hand scale is mres = ( 2GF ne 2Eν )1/2
which indicates the resonance value for (∆m2ν )1/2 for Eν = 10 MeV, assuming
an electron number fraction of Ye = 0.5.
sin2 2θ > −8
∼ 10 Eν /10 MeV. (11.14)
capture. As discussed in Sect. 11.1.4, the hot bubble between the set-
tled protoneutron star and the escaping shock wave at a few seconds
after core bounce might be an ideal high-entropy environment for this
process for which no other site is currently known that could reproduce
the observed galactic heavy element abundance and isotope distribu-
tion. Naturally, the r-process can only occur in a neutron-rich medium
(Ye < 12 ). The p/n ratio in the hot bubble is governed by the β reac-
tions νe n ↔ pe− and ν e p ↔ ne+ . Because the neutrino number density
is much larger than the ambient e+ e− population the proton/neutron
fraction is governed by the neutrino spectra and fluxes. The system
is driven to a neutron-rich phase because normally the ν e ’s are more
energetic than the νe ’s. They emerge from deeper and hotter regions of
the star because their opacity is governed by the same reactions, and
because the core is neutron rich, yielding a larger opacity for νe .
If an exchange νe ↔ νµ,τ occurs outside of the neutrino sphere the
subsequent νe flux is more energetic than the ν e flux which did not
undergo a swap. (A normal mass hierarchy has been assumed.) Even
a partial swap of a few 10% is enough to shift the medium to a proton-
rich state, i.e. to Ye > 12 , to be compared with the standard values of
0.35−0.46. Therefore, the occurrence of such oscillations would be in
conflict with r-process nucleosynthesis in supernovae (Qian et al. 1993).
Fig. 11.20. Density profile for a SN model at 6 s after the core bounce,
typical for √
the “hot bubble phase” (Qian et al. 1993). The right-hand scale
is mres = ( 2GF ne 2Eν )1/2 which indicates the resonance value for (∆m2ν )1/2
for Eν = 10 MeV, assuming Ye = 0.5. (Note that out to a few km above the
neutrino sphere Ye . 0.5.)
Supernova Neutrinos 439
Fig. 11.21. Mass difference and mixing angle of νe with νµ or ντ where a spec-
tral swap would be efficient enough to help explode supernovae (schemat-
ically after Fuller et al. 1992), and where it would prevent r-process nu-
cleosynthesis (schematically after Qian et al. 1993; Qian and Fuller 1994).
440 Chapter 11
be taken to be the same at 3×1051 erg s−1 each, and R = 11 km. For
these conditions the estimated contributions to ∆V from electrons and
neutrinos are shown in Fig. 11.22 as a function of radius.
As long as no swap has occurred the neutrinos do not play a major
role because the contribution of νµ cancels exactly against ν µ . Further,
the difference between νe and ν e is smaller than the electron contribu-
tion, and it has the same sign whence it simply causes a slightly larger
effective matter density.
However, on resonance a relatively large number of νe ’s exchange
flavor with νµ ’s so that the neutrino contribution changes sign. More-
over, on resonance the vacuum ∆m2ν by definition cancels against the
medium-induced contribution. Therefore, “switching on” the neutrino
term shifts the resonance position for given vacuum mixing parameters.
Also, in a self-consistent treatment one needs to consider the full non-
linear equations of motion for the neutrino density matrix which causes
an “off-diagonal refractive index” as discussed in Sect. 9.3.2. Qian and
Fuller (1995) have performed an approximately self-consistent analysis
of this problem. All told, they found that the neutrino-neutrino interac-
tions have a relatively small impact on the parameter space where flavor
conversion disturbs the r-process. Within the overall precision of these
arguments and calculations, their final exclusion plot is nearly identi-
cal with the schematic picture shown in Fig. 11.21. Qian and Fuller’s
findings are corroborated by a study performed by Sigl (1995a).
c) Summary
In summary, there remains a large range of mixing angles where neu-
trino oscillations between νe and νµ or ντ with a cosmologically in-
teresting mass could help to explode supernovae, and yet not disturb
r-process nucleosynthesis. If one assumes a mass hierarchy with νe
dominated by the lightest, ντ by the heaviest mass eigenstate, the cos-
mologically relevant neutrino would be identified with ντ . It is inter-
esting that the relevant range of mixing angles, 3×10−4 < <
∼ θ ∼ 3×10 ,
−3
overlaps with the mixing angle among the first and third family quarks
which is in the range 0.002−0.005 (Eq. 7.6). Therefore, a scenario
where a massive ντ plays a cosmologically important role, helps to ex-
plode supernovae, and leaves r-process nucleosynthesis unscathed does
not appear to be entirely far-fetched. This scenario leaves the possi-
bility open that the MSW effect solves the solar neutrino problem by
νe -νµ oscillations.
11.4.6 A Caveat
All existing discussions of neutrino flavor oscillations in SNe were based
on spherically symmetric, smooth density profiles. However, there
can be significant density variations, convection, turbulence, and so
forth. Therefore, it is clear that many of my statements about medium-
induced oscillation effects are provisional. Further studies will be re-
quired to develop a more complete picture of SNe and their neutrino
oscillations as 3-dimensional events.
A first study of SN neutrino oscillations with an inhomogeneous den-
sity profile was recently performed by Loreti et al. (1995). They added
a random density field to the standard smooth profile and studied the
impact on neutrino oscillations. They found that the shock-revival sce-
nario involving MSW oscillations can be significantly affected in the
Supernova Neutrinos 443
12.1 Preliminaries
This book is largely about the properties of electrically neutral particles
whose electromagnetic interactions are correspondingly weak. However,
because they can virtually dissociate into charged states, they will still
interact with photons through higher-order amplitudes. The focus of
the present chapter is the possibility of radiative decays of the form
ν → ν ! γ (neutrinos) or a → γγ (axions), but also ν → ν ! e+ e− and
ν → ν ! e+ e− γ. Cowsik (1977) was the first to recognize that the huge
path lengths available in the astrophysical environment allow one to
obtain much more restrictive limits on such decays than from labora-
tory experiments. For example, the absence of single-photon counts in
a detector near a fission reactor indicates a bound69 τγ /mνe > 22 s/eV
69
In this chapter τγ will always denote the partial neutrino decay time into radia-
tion while τtot is the total decay time if hypothetical invisible channels are included.
−1
Then τγ−1 = Bγ τtot with the branching ratio Bγ .
449
450 Chapter 12
one can then derive limits on Ueh for any νh , even a hypothetical sterile
one, if mh >
∼ 1 MeV because it is the same mixing amplitude that allows
for its production in the source and for its νh → νe e− e+ decay.
In the following I will discuss radiative lifetime limits from different
sources approximately in the order of available decay paths, from labo-
ratory experiments (a few meters) to the radius of the visible universe
(about 1010 light years).
where ϑ is the angle between the ν spin polarization vector and the
photon momentum. For Majorana neutrinos the decay is isotropic,
independently of their polarization, and thus α = 0. Similarly for
axion decays, a → γγ, as these particles have no spin so that in their
rest frame no spatial direction is favored. For polarized Dirac neutrinos
the possible parameter range is −1 ≤ α ≤ 1.
The decay photon has an energy ω = δm mν /2 in the rest frame of
the parent neutrino—see Eq. (7.12). If it is emitted in a direction θ
with regard to the laboratory direction of motion, the energy in the
laboratory frame is Eγ = ω (Eν + pν cos θ)/mν . Hence,
where β = pν /Eν = (1 − m2ν /Eν2 )1/2 is the neutrino velocity. For a left-
handed parent neutrino the spin is polarized opposite to its momentum
so that cos θ = − cos ϑ and dNγ /d cos θ = 12 (1 + α cos θ). For a very
452 Chapter 12
Fig. 12.1. Photon spectrum from the decay of a relativistic neutrino (en-
ergy Eν ) according to Eq. (12.4).
A fraction (mν /Eν )(dγ /τγ ) of neutrinos decay before reaching the
detector if the laboratory lifetime is large compared with the decay
path dγ . Here, τγ is the rest-frame radiative decay time and Eν /mν
the time dilation factor. Integrating over the neutrino source spectrum
then yields
! "
mν # ∞ dEν Eγ Fν (Eν )
Fγ (Eγ ) = dγ 1 − α + 2α . (12.5)
τγ Eγ /δm δm δm E ν Eν2
Fig. 12.2. Spectrum of reactor ν e ’s per fission of 239 Pu and 235 U (von Feil-
itzsch et al. 1982; Schreckenbach et al. 1985). The width of the lines gives
the total error of the spectra.
able bound was inferred by Vogel (1984) from data taken at the Gösgen
reactor (Switzerland). The most recent analysis is, again, based on
data taken with a scintillation counter at Gösgen (Oberauer, von Feil-
itzsch, and Mössbauer 1987). From a comparison of the “reactor on”
with the “reactor off” photon counts for several energy channels in the
MeV range and using the experimentally established neutrino spectrum
(Fig. 12.2), these authors found the 68% CL lower limits on the ν e ra-
diative decay times of τγ /mνe > 22 s/eV for α = −1, 38 s/eV for α = 0,
and 59 s/eV for α = +1. It was assumed that ν ! is massless so that
δm = 1 in Eq. (12.5). With Eq. (7.12) the α = −1 constraint translates
into a bound on the effective electromagnetic transition moment of
not a very restrictive limit even if νe saturates its upper mass bound of
about 5 eV.
Even this weak limit would cease to apply if νe and ν ! became
nearly degenerate. Therefore, Bouchez et al. (1988) performed an ex-
periment where they searched for optical decay photons at the Bugey
reactor (France). They excluded a certain region in the parameter
plane spanned by τγ /mνe and δm . Their greatest sensitivity was ap-
proximately at δm = 2×10−5 where they found τγ /mνe > ∼ 0.04 s/eV.
< 7 2
With Eq. (7.12) this is µeff ∼ 10 µB /meV which, unfortunately, is ir-
relevant as a constraint. Hence, for nearly degenerate neutrinos there
Radiative Particle Decays 455
With mh up to an MeV this bound has a lot more teeth than the one
on electron neutrinos.
For mh > 2me ≈ 1 MeV the decays νh → νe e+ e− will become
kinematically possible and probably dominate. The scintillation coun-
ters that were used to search for decay photons near a power reactor
are equally sensitive to electrons and positrons—for many purposes
relativistic charged particles may be treated almost on the same foot-
ing as γ rays. Therefore, the same Gösgen data have been analyzed
to constrain the mixing amplitude Ueh (Oberauer, von Feilitzsch, and
Mössbauer 1987; Oberauer 1992). Even more restrictive limits were
obtained from data taken at the Rovno reactor (Fayons, Kopeykin,
and Mikaelyan 1991), and most recently at the Bugey reactor (Hagner
et al. 1995). Note that in this method the same mixing probability
|Ueh |2 appears in Eq. (12.1) to obtain the νh flux from a νe source, and
in Eq. (7.9) to obtain the decay probability. Hence, the expected e+ e−
flux is proportional to |Ueh |4 .
456 Chapter 12
For the admixture of other mass eigenstates this result may be trans-
lated in a fashion analogous to the discussion of reactor neutrinos.
Beam-stop neutrinos from meson decays may also be used to con-
strain the e+ e− decays of heavy admixtures. Because of the larger
amount of available energy one may probe higher masses for νh while
the reactor bounds drop out above a few MeV because of the relatively
soft spectrum. In Fig. 12.3 the most restrictive such constraints are
summarized.
71
It is customary to display |Ueh |2 when constraining the mixing parameters of
heavy, decaying neutrinos while one shows sin2 2θeh for light, oscillating neutrinos
as in Chapter 8. For easier comparison I always use the mixing angle. Recall that
|Ueh | = sin θeh so that for small mixing angles sin2 2θeh = 4|Ueh |2 .
Radiative Particle Decays 457
Like a terrestrial power reactor, the Sun is a prolific neutrino source ex-
cept that it emits νe ’s rather than ν e ’s. The expected spectrum as well
as the relevant measurements were discussed in Chapter 10. Suffice
it to recall that the solar neutrino flux is now experimentally estab-
lished without a shred of doubt. There remain significant discrepancies
between the predicted and measured spectral shape of the spectrum
which may be explained by neutrino oscillations. However, the solar
neutrino problem is a fine point in the context of the present discus-
sion because the following results depend mostly on the low-energy
pp flux.
One may proceed exactly as in the previous section in order to trans-
late the solar neutrino spectrum shown in Fig. 10.1 into an expected
flux of x- and γ-rays from the Sun. In Fig. 12.4 I show this flux at Earth
for τγ /mνe = 10 s/eV (about the laboratory lifetime limit) and for the
values ±1 for the “anisotropy parameter” α. The spectrum as shown is
based on the calculated neutrino flux. The shoulders corresponding to
other than the pp neutrinos likely would have to be reduced somewhat.
The magnitude of the photon flux is enormous because the decay path
dγ is the entire distance to the Sun of 1.5×1013 cm = 500 s.
The quiet Sun is a significant source of soft x-rays from the quasi-
thermal emission of the hot corona at T ≈ 4.5×106 K. The flux mea-
surements of Chodil et al. (1965) are marked as open diamonds in
Fig. 12.4. The flux of decay photons in Fig. 12.4 would outshine the
solar corona by some 4 orders of magnitude! Moreover, the corona
spectrum falls off sharply at larger energies. In the hard x- and soft
γ-ray band very restrictive upper limits exist on the emission of the
quiet Sun that are shown in Fig. 12.4. They are based on balloon-
borne detectors flown many years ago (Frost et al. 1966; Peterson et al.
1966). These upper limits are still far above the estimated albedo (ra-
diation from cosmic rays hitting the surface of the Sun), leaving much
room for improvement. Alas, the quiet Sun is not an object of great
interest to γ-ray astronomers and so more recent measurements do not
seem to exist.
In order to respect these measured upper limit photon fluxes one
must shift the decay spectrum in Fig. 12.4 down by about 8 orders of
magnitude (thin solid line in Fig. 12.4). This yields a lower radiative
liftime limit for νe of τγ /mνe > 9
∼ 7×10 s/eV (Cowsik 1977; Raffelt 1985).
Radiative Particle Decays 459
Fig. 12.4. Spectrum of photons from the solar neutrino decay νe → ν ! γ for
the indicated values of the anisotropy parameter α. Measurements of the
x-ray emission of the solar corona (open diamonds) according to Chodil et
al. (1965). Upper limit x- and γ-ray fluxes according to Frost et al. (1966)
and Peterson et al. (1966). Estimated albedo according to Peterson et al.
(1966). Thin solid line: Maximally allowed photon spectrum from neutrino
decay. (Figure adapted from Raffelt 1985.)
Fig. 12.5. Excluded parameters for νe → ν ! γ from the Sun for nearly degen-
erate neutrino masses (adapted from Raffelt 1985).
Because they used the theoretically expected rather than the experi-
mentally measured solar 8 B neutrino flux I have discounted their orig-
inal numbers by a factor of 3. These bounds are included in Fig. 12.3;
in the applicable mass range they are more restrictive than those from
laboratory experiments. They are valid only if the νh flux is not dimin-
ished by invisible decay channels on its way between Sun and Earth, i.e.
the total νh (laboratory) lifetime must exceed about 500 s. Because in
the mass range of a few MeV the time dilation factor for about 10 MeV
neutrinos is not large, the bounds apply for total νh lifetimes exceeding
about 100 s. Such “long-lived” MeV-mass neutrinos are in conflict with
the big-bang nucleosynthesis constraints shown in Fig. 7.2.
462 Chapter 12
Besides neutrinos, stars can also produce other weakly interacting par-
ticles by both plasma and nuclear processes. With a temperature of
about 1.3 keV in the solar center the former reactions would produce a
relatively soft spectrum and so I focus on nuclear reactions where MeV
energies are available. If the new particle is a boson and if it couples
to nucleons, it will substitute for a photon with certain relative rates r
in reactions with final-state γ-rays. A short glance at the nuclear reac-
tion chains shown in Fig. 10.2 reveals that a particularly useful case is
p + d → 3 He + γ with Eγ = 5.5 MeV. This reaction occurs about 1.87
times for every 4 He nucleus produced by fusion in the Sun and so it
must occur about 1.7×1038 s−1 . A certain fraction of the particles pro-
duced will be reabsorbed or decay within the Sun. If their probability
for escaping is p the Sun emits r p 1.7×1038 s−1 of the new objects.
If the new particle is a scalar boson a it will have a decay channel
a → 2γ. Because this decay is isotropic in a’s rest frame the spectrum
of decay photons is box-shaped (Fig. 12.1) with an upper endpoint of
5.5 MeV. If a fraction q of the particles decays between the Sun and
Earth the local γ flux is r p q 2.2×1010 cm−2 s−1 MeV−1 . The upper limit
photon flux shown in Fig. 12.4 at 5.5 MeV is 0.8×10−3 cm−2 s−1 MeV−1
(Peterson et al. 1966). From there, Raffelt and Stodolsky (1982) found
the general upper bound r p q < 4×10−14 . They also calculated r, p,
and q for the specific case of “standard axions” and were able to derive
a strong limit on the properties of this hypothetical particle. Together
with many laboratory constraints (Particle Data Group 1994) standard
axions are now entirely excluded, the main motivation to consider “in-
visible axions” instead (Chapter 14).
emission at Tνe ≈ 4 MeV so that the fluence73 of ν’s plus ν’s per flavor
was about
mν dLMC ( )
Fγ! (Eγ ) = Fν (1 − α) e −ε
+ 2α ε E 1 (ε) , (12.15)
τγ 2Tν2
Fig. 12.7. Event rates measured in the Gamma Ray Spectrometer (GRS)
of the Solar Maximum Mission (SMM) satellite encompassing the observed
neutrino burst of SN 1987A (the dashed line is for the first neutrinos observed
in the IMB detector). The time interval for each bin is 2.048 s. The rates to
the left of the dashed line are used to determine the background while the
ones to the right would include photons from neutrino decay. (Figure from
Oberauer et al. 1993 with permission.)
For the 10 s and 223.2 s time intervals one can compute an average
flux limit (cm−2 s−1 ) for each channel. Then one expects that for the
longer time interval it is more restrictive by the ratio of (∆t)1/2 , i.e. by
(10 s/223.2 s)1/2 = 0.21. This expectation is approximately borne out
by the data in Tab. 12.1, confirming their consistency.
Radiative Particle Decays 467
Therefore, using the same Boltzmann source spectrum the photon spec-
trum is slightly harder. Going through the same steps as before one
finds the upper limits on Bγ as a function of the assumed Tν and α
75
Somewhat stronger constraints found in the literature were based on a less de-
tailed analysis, notably with regard to the spectral dependence and the dependence
on α. Published results are τγ /mν > 0.83×1015 s/eV (von Feilitzsch and Oberauer
1988), 1.7×1015 (Kolb and Turner 1989), 6.3×1015 (Chupp, Vestrand, and Reppin
1989), and 2.8×1015 (Bludman 1992).
468 Chapter 12
Fig. 12.9. Upper limit on the radiative branching ratio Bγ for mν <
∼ 40 eV
< 5
and τtot /mν ∼ 5×10 s/eV (most neutrinos decay between SN 1987A and
Earth).
where τs = τtot /s. Of course, if τtot became so short that the neutrinos
would decay while still within the envelope of the progenitor even this
Radiative Particle Decays 469
weak limit would not apply. This is the case for τlab < ∼ Renv ≈ 100 s
(envelope radius Renv of the progenitor star) and so with Eν ≈ 20 MeV
one needs to require τtot /mν > −5
∼ 10 s/eV.
The photons which arrive first are the ones from decays near the
source. Then the limit dD ( dLMC leads to dγ = dLMC − dD cos θlab and
t = tD (1 − β cos θlab ). With cos θlab = (β + cos θ)/(1 + β cos θ) where θ
is the angle of photon emission in the parent frame one finds
tD
t= , (12.20)
γ 2 (1 + βx)
where γ = Eν /mν is the neutrino Lorentz factor and x ≡ cos θ. There-
fore, photons detected between t and t + dt result from decays at
tD = γ 2 (1 + βx) t during an interval dtD = γ 2 (1 + βx) dt. (Recall
that t is measured after the first massless neutrinos arrived while tD
is measured after emission at the source.) The number of parent neu-
trinos diminishes in time as e−tD /γτtot with τtot the total decay time.
Therefore, the number of photons traversing a spherical shell of radius
dLMC per unit time is
γ(1 + βx) −γ(1+βx) t/τtot
Ṅγ (t) = e , (12.21)
τγ
where as before τγ is the radiative decay time.
Photons produced before the parent has left the envelope of the
progenitor star (radius Renv ) cannot be detected at Earth. If this
absorption effect is to be included, Eq. (12.21) will involve a step
function76 Θ(dD − Renv ) with dD = βtD = βγ 2 (1 + βx) t. Put an-
other way, photons emitted at an angle θ in the rest frame will first
arrive at a time t0 = Renv [βγ 2 (1 + βx)]−1 . The progenitor of SN 1987A
has been unambiguously identified as the blue supergiant Sanduleak
−69 202 (Schramm and Truran 1990). From its surface temperature
(15,000 K), its luminosity (5×1038 erg/s) and the distance to the LMC
one can infer its radius to be Renv ≈ 3×1012 cm = 100 s. Using this
76
The step function is defined by Θ(z) = 0 for z < 0 and Θ(z) = 1 for z > 0.
Radiative Particle Decays 471
Fig. 12.11. Arrival time t0 of first decay photons from a parent neutrino
with velocity β, taking the envelope radius of the source to be Renv = 100 s.
The curves are marked with the respective values of x = cos θ, the direction
of photon emission in the neutrino rest frame.
where ω± = Eγ [γ(1 ± β)]−1 and x = (Eγ /γω − 1)/β. The photon flux
at Earth is obtained by multiplication with the neutrino fluence Fν of
Eq. (12.12) and integration over a suitable spectrum Φν (Eν ) of neutrino
energies.
If the neutrinos are sufficiently long-lived (the exact meaning of
this is quantified below) the exponential can be ignored. If one also
ignores the absorption effect by the progenitor star (Renv = 0), the
472 Chapter 12
time structure of Eq. (12.22) reduces to Θ(t), i.e. its spectral form is
time independent except that it begins at t = 0. This is somewhat
surprising because the energy of a photon in the laboratory frame is
related to the angle of emission in the neutrino rest frame which in
turn determines the “detour” taken from the source to us (Fig. 12.10).
However, the first photons come from decays immediately at the source
and so any angle of emission leads to the same initial arrival time.
It must be stressed that the expression Eq. (12.22) depends on the
assumption of decays not too far from the source (dD ( dLMC ) and so
only the head of the photon pulse is correctly described while its tail
would require including decays even close to the Earth. Strictly speak-
ing, the photon burst never ends because even if the parent neutrinos
have passed the Earth, some photons will be received from backward
emission. However, because one is interested in neutrino masses so
large (mν >∼ 200 eV) that the photon burst is much longer than the
GRS measurement window, it is enough to account for the head of the
photon pulse.
ignored (tenv = 0) the spectrum is shown in Fig. 12.12 for the anisotropy
parameters α = 0, ±1. Note the difference to the triangular shape of
Fig. 12.1 for a stationary source. High-energy photons are now en-
hanced because lower-energy ones correspond to larger emission angles
in the parent frame and so they take a larger “detour” from the source
to us (Fig. 12.10). Hence, their flux is spread out over a larger time
interval even though photons of all energies begin to arrive at the same
time if tenv = 0.
Fig. 12.12. Photon spectrum from the decay of a short burst of relativistic
neutrinos, energy Eν , according to Eq. (12.24) taking pν = Eν and e−t/τ∗ = 1
(relativistic and long-lived parent), and ignoring absorption effects by the
progenitor (tenv = 0).
In order to compare with the GRS fluence limits one needs to in-
tegrate the expected flux between t = 0 and t = tGRS = 223.2 s.
The Θ function in Eq. (12.24) is accounted for by using tenv as a lower
limit of integration. Integrating also over the neutrino source spectrum
Fν Φν (Eν ) yields
0 1
tGRS # ∞ 2Eγ 2Eγ − Eν
Fγ! = Fν dEν Φν 1+α I, (12.27)
mν τγ Emin pν pν
where
e−tenv /τ∗ − e−tGRS /τ∗
I≡ . (12.28)
tGRS /τ∗
For sufficiently long-lived parents (τ∗ , tGRS ) the exponentials can
be expanded and I = (tGRS − tenv )/tGRS . The lower limit of integra-
tion is set by the condition Eq. (12.26) and by the requirement that
tenv < tGRS , i.e. that I > 0. This condition may be expressed as
pν > (m2ν /2Eγ ) (Renv /tGRS ).
474 Chapter 12
1 − e−tGRS /τ∗
I= , (12.29)
tGRS /τ∗
Fig. 12.15. (The overall factor m−1 ν of Eq. 12.32 is not included, of
course.) Up to neutrino masses of about 10 MeV one may essentially
ignore the nonrelativistic corrections while for larger masses one has to
worry about them. However, because this discussion applies to stan-
dard neutrinos, the largest relevant mass is about 24 MeV and so the
nonrelativistic corrections never overwhelm the result.
Fig. 12.16. Different regimes of neutrino masses and total lifetimes referred
to in the text. ∆tγ is the duration of the burst of decay photons. The
radiative lifetime limits in the areas 1−6 are summarized in Tab. 12.2.
−1/2
5 mν < 107 Bγ < 3×10−10 0.7×10−5 m−3/2
ν τtot
τtot mν < 4×108
10−5 < τtot m−1
ν < 3×10
5
6 τtot m−1
ν < 10
−5
Bγ < 0.01 —
a Forthe anisotropy parameter α = −1.
b Numbered as in Fig. 12.16.
c Neutrino masses in eV, lifetimes in s.
d Upper limit.
trinos have Dirac masses and thus right-handed partners which could
be emitted from the inner core of the SN by helicity-flipping processes
(Sect. 13.8). Moreover, entirely new particles could be produced and
escape from there.
The present bounds can be scaled to such cases if one calculates the
total energy Ex,tot = fx 1053 erg emitted in the new x particles, where
1×1053 erg is the total energy that was used for a standard ν plus ν.
Self-consistency requires fx < 1, of course. In addition, one needs the
average energy )Ex * of the new objects which allows one to define an
approximate equivalent temperature Tx = 13 )Ex *. Depending on the x
mass and total lifetime one can then read the radiative lifetime limits
directly from Figs. 12.8, 12.9, and 12.14, except that they must be
relaxed by a factor fx for the reduced fluence.
480 Chapter 12
12.4.7 Limit on ντ → νe e+ e−
In order to estimate the expected photon flux from Eq. (12.22) one
needs to know the distribution of photon energies and emission angles
in the frame of the parent neutrino. In the absence of a detailed calcu-
lation I follow Oberauer et al. (1993) and assume approximate isotropy
for the photon emission. The soft part of the spectrum dNγ /dω from
a bremsstrahlung process is given by the rate of the primary process
times (α/π) ω −1 (Jackson 1975). Extending this behavior up to photon
energies of 12 mν I use
1 α/π 1
f (ω, x) = Θ( 12 mν − ω) , (12.34)
τγ τe+ e− 2ω
an expression which does not depend on x because of the assumed
isotropy. (α = 1/137 is the fine-structure constant, not the previous
anisotropy parameter.)
After dropping the exponential in Eq. (12.22) because the neutri-
nos are long-lived, one integrates over a Boltzmann source spectrum
for nonrelativistic neutrinos, integrates over the GRS energy channels,
and compares the expected fluence with the measured upper limits of
Tab. 12.1. The resulting bound on τe+ e− corresponds with Eq. (7.9)
directly to a limit on |Ue3 |2 . In Fig. 12.18 I show these bounds (trans-
formed into bounds on the mixing angle) as a function of the assumed
neutrino mass for Tν = 4 and 6 MeV. There remains a strong limit
even for masses far exceeding the temperature because the exponential
The bounds on Ue3 from reactors, beam stops, or the Sun were based on
a νe flux which partially converts into ν3 ’s which subsequently decay.
Because the SN emits about equal numbers of all ordinary neutrino
flavors, this approach is obsolete with regard to ν3 . However, one may
still consider hypothetical sterile neutrinos which interact only by virtue
of their mixing with νe . By assumption these states would not interact
through ordinary weak interactions and so they would not be trapped
in the SN core. Hence the expected νh flux would emerge from the deep
interior rather than the surface of the core.
In this case, however, the sterile neutrinos would carry away energy
much more efficiently than the ordinary ones and so the requirement
that enough energy was left for the observed ν e ’s from SN 1987A al-
ready gives one the approximate limit |Ueh |2 < ∼ 10
−10
(Sect. 9.6). If
this limit is approximately saturated one expects that about as much
energy is carried away by νh as by the ordinary flavors. Taking account
of the harder energies of neutrinos emitted from the SN core one still
obtains about the same limit on |Ueh |2 as on |Ue3 |2 before. Put an-
other way, the GRS observations do not dramatically improve on the
cooling argument of Sect. 9.6, although they range in the same general
magnitude.
Radiative Particle Decays 483
12.4.9 Axions
Another hypothetical particle that could have been emitted abundantly
from SN 1987A is the axion. The impact of the axionic energy loss is
discussed in Sect. 13.5. If axions are more strongly interacting than a
certain limit, implying that their mass is larger than a few eV, they
are emitted from the surface of the SN core with a luminosity similar
to that of neutrinos. Kolb and Turner (1989) found that the axion
−12/11
fluence from SN 1987A would have been Fa ≈ 6×1010 cm−2 meV
−4/11
with meV ≡ ma /eV at a temperature of Ta ≈ 15 MeV meV . From
the GRS fluence limits, Kolb and Turner found that ma must be less
than a few 10 eV, a bound which is less restrictive than, for example,
the limit from globular cluster stars (Sect. 5.2.5).
Fig. 12.19. Excluded areas of the ντ mass and lifetime if the standard-model
decay ντ → νe e+ e− is the only available channel. The laboratory results
refer to the bounds on sin2 2θe3 of Fig. 12.3, translated into a limit on τe+ e−
by virtue of Eq. (7.9). The SN 1987A bound is that from Fig. 12.18 while
the cosmological one is from Fig. 7.2. The excluded range indicated by the
vertical arrow refers to the argument of Sect. 12.5.1.
d 2 Fγ mν 1 9 Ṅν t2U
= , (12.35)
dEγ dΩ τγ 4π 21/2 5 Eν3/2 Eγ1/2
where tU is the age of the universe. Moreover, it was assumed that
in ν → ν ! γ the daughter neutrino is massless, and that the decays
are isotropic in the parent frame (anisotropy parameter α = 0). In
a flat universe one has tU = 23 H0−1 = h−1 2.05×1017 s where H0 =
Radiative Particle Decays 487
on the Hubble type, varying from 0.2 h2 SNu for Sa spirals to about
5 h2 SNu for Sd (van den Bergh and Tammann 1991) where the su-
pernova unit is defined by 1 SNu ≡ 1 SN per century per 1010 L%,B .
Adopting 1 h2 SNu as a representative value and about 5×1057 neu-
trinos plus antineutrinos of a given flavor per SN yields for each fla-
vor Ṅν ≈ h3 1.3×10−27 cm−3 s−1 , a rate almost identical to that from
hydrogen-burning stars.79 The radiative lifetime limit is then also iden-
tical, except that it applies to neutrinos of all flavors.
All of these bounds are weaker than those from SN 1987A. There-
fore, decaying stellar neutrinos cannot actually contribute to the ob-
served x- and γ-ray background.
Fig. 12.21. Limits on µeff according to Ressell and Turner’s (1990) bounds
on the radiative lifetime of long-lived neutrinos from the diffuse cosmic back-
ground radiations. The dashed line corresponds to Eq. (12.38).
still play a significant dynamical role. Recently, such mixed dark matter
scenarios have received much attention where mν = 5 eV is a favored
value. Such low-mass particles cannot cluster on galactic scales, but
likely they would reside in clusters of galaxies. With radiative decays
ν → ν ! γ and a total lifetime exceeding the age of the universe one then
expects clusters of galaxies to be strong sources of optical or ultraviolet
photons. Several limits are summarized in Fig. 12.22.
A case has been made that radiatively decaying neutrino dark mat-
ter is actually required to solve certain problems, notably the ioniza-
tion of galactic hydrogen clouds (e.g. Melott and Sciama 1981; Melott,
McKay, and Ralston 1988; Sciama 1990a,b; Sciama 1993a,b, 1995).
The predictions are very specific: An energy of decay photons of Eγ =
(14.4 ± 0.5) eV and thus a neutrino mass of mν = (28.9 ± 1.1) eV
with a radiative lifetime of τγ = (2 ± 1) × 1023 s which translates into
µeff = (6.3 ± 2) × 10−15 µB . Such a large transition moment would
require particle physics beyond the standard model.
Nominally, this possibility is already excluded by the absence of a
uv line from the cluster A665 (Davidsen et al. 1991). However, a bound
from a single source is always subject to the uncertainty of unrecognized
absorbing material in the line of sight or internal absorption. Moreover,
the dark matter in the core of this cluster may be mainly baryonic
(Sciama, Persic, and Salucci 1993; Melott et al. 1994). Bounds from the
diffuse extragalactic background light are more reliable in this regard.
While they marginally exclude Sciama’s neutrino (Overduin, Wesson,
and Bowyer 1993) it is perhaps too early to pronounce it entirely dead.
A decisive test will be performed with a future satellite experiment
where the uv line from neutrinos decaying in the solar neighborhood
definitely would have to show up if neutrinos were the bulk of the
galactic dark matter (e.g. Sciama 1993b).
12.7.2 Axions
The axion lifetime from a → 2γ is τ = 6.3×1024 s (ma /eV)5 /ξ 2 where ξ
is a model-dependent number of order unity (Sect. 14.3.2). Therefore,
axions with eV masses have radiative lifetimes in the neighborhood of
the above neutrino limits whence they can be constrained by similar
methods (Kephart and Weiler 1987). Moreover, such axions would con-
tribute substantially to the mass density of the universe because they
would have been in thermal equilibrium until relatively late.80 Their
80
In the early universe, axions are also produced by the relaxation of the coherent
initial field configuration at the onset of the QCD phase transition. This process
492 Chapter 12
The most interesting limits arise from a search for axion decay lines
from the intergalactic space in the clusters of galaxies A2256 and A2218.
Bershady, Ressell, and Turner (1991) and Ressell (1991) found ξ < 0.16
0.078, 0.039, 0.032, 0.016, and 0.011 for ma /eV = 3.5, 4.0, 4.5, 5.0,
6.0, and 7.5 respectively (Fig. 12.23). These limits are placed into the
context of other constraints in Fig. 5.9.
yields Ωa h2 ≈ (10−5 eV/ma )1.175 —see Eq. (14.5). While the overall coefficient of
this expression is very uncertain it is clear that Ωa = 1 saturates for ma somewhere
between 1 µeV and 1 meV. In this range axions never achieved thermal equilibrium.
Chapter 13
The lessons for particle physics from the SN 1987A neutrino burst are
studied. First, neutrinos could have decayed or oscillated into other
states on their way out of the SN core and to us. Second, propagation
effects could have caused a time delay between photons and neutrinos
or between ν e ’s of different energy. Third, nonstandard cooling agents
could have shortened the neutrino burst below its observed duration.
These arguments are applied to a variety of specific cases.
13.1 Introduction
In Chapter 11 the neutrino observations from SN 1987A were dis-
cussed and it was shown that they agree well with standard theoret-
ical expectations from the core collapse and subsequent explosion of
an evolved massive star. The signal displays several anomalies (time
gap at Kamiokande, anisotropy in both detectors) which render it a
less beautiful specimen of the expected signal characteristics than is
sometimes stated in the literature. Still, in the absence of plausible
alternatives one must accept that the Kamiokande II, IMB, and Bak-
san event clusters observed at 7:35 UT on 27 February 1987 represent
the ν e component of the neutrino burst from the core collapse of the
SN 1987A progenitor star rather than some other particle flux, or some
other reaction than the expected dominant ν e p → ne+ process.
Accepting this, there is a host of consequences concerning a variety
of fundamental physics issues. The first and simplest set of arguments
is based on the fact that the ν e pulse and perhaps the prompt νe burst
493
494 Chapter 13
However, this simple result must be interpreted with care because mas-
sive neutrinos are expected to mix. The heavy νe admixtures could
decay and may violate this bound.
The ν e ’s were not removed by excessive scattering on cosmic back-
ground neutrinos, majorons, dark-matter particles etc., leading to con-
straints on “secret interactions” (Kolb and Turner 1987). Take the
What Have We Learned from SN 1987A? 495
so that the speed of light and that of neutrinos are equal to within
(Longo 1987; Stodolsky 1988)
! !
!c − c !
! ν γ! <
! ! 2×10−9 , (13.3)
! cγ ! ∼
where the integral is taken along the trajectory r(t) of the beam between
the points of emission (E) and absorption (A). This delay is the same
for neutrinos and photons to within
! !
! ∆t − ∆t !
! ν γ!
! ! < 0.7−4×10−3 , (13.5)
! ∆tγ !
b) Neutrino Charge
The absence of an energy-dependent dispersion of the neutrino pulse
can be used to constrain other neutrino properties. A small electric
charge eν would bend the neutrino path in the galactic magnetic field,
leading to a time delay of
∆t e2 (BT dB )2
= ν , (13.6)
t 6Eν2
where BT is the transverse magnetic field and dB the path length within
the field. This leads to a constraint of
# $# $
eν < 1 µG 1 kpc
∼ 3×10−17 (13.7)
e BT dB
(Barbiellini and Cocconi 1987; Bahcall 1989). Note that a typical field
strength for the ordered magnetic field in the galactic spiral arms is
2−3 µG and that the path length of the neutrinos within the galactic
disk is only of order 1 kpc because the LMC lies high above the disk
(galactic latitude about 33◦ ).
84
Limits on mνe from the SN 1987A data were derived, among others, by Abbott,
de Rújula, and Walker (1988), Adams (1988), Arnett and Rosner (1987), Bah-
call and Glashow (1987), Burrows and Lattimer (1987), Burrows (1988a), Chiu,
Chan, and Kondo (1988), Cowsik (1988), Kolb, Stebbins, and Turner (1987a,b),
Midorikawa, Terazawa, and Akama (1987), Sato and Suzuki (1987a,b), Spergel and
Bahcall (1988), Loredo and Lamb (1989), and Kernan and Krauss (1995).
500 Chapter 13
c) Long-Range Forces
Speculating further one may imagine some sort of neutrino “fifth-force
charge.” If electrons, protons, or dark-matter particles also carry such
a charge the bending of the neutrino trajectory in the fifth-force field of
the galaxy would lead to an energy-dependent time delay. This and re-
lated arguments were advanced by a number of authors (Pakvasa, Sim-
mons, and Weiler 1989; Grifols, Massó, and Peris 1988, 1994; Fiorentini
and Mezzorani 1989; Malaney, Starkman, and Tremaine 1995).
The most plausible form for such a long-range interaction is one me-
diated by a massless vector boson, i.e. a new gauge interaction, perhaps
related to a novel leptonic charge (Sect. 3.6.4). In this case neutrinos
and antineutrinos would carry opposite charges so that the cosmic neu-
trino background would be essentially a neutral plasma with regard to
the new interaction. The resulting screening effects then invalidate the
SN 1987A argument (Dolgov and Raffelt 1995).
Screening effects would not operate if the force were due to a spin-0
or spin-2 boson which always cause attractive forces. However, any
force mediated by a massless spin-2 boson must couple to the energy-
momentum tensor and thus is identical with gravity. The force medi-
ated by a scalar boson between a static source and a relativistic neutrino
is suppressed by a Lorentz factor. Therefore, even if scalar-mediated
forces existed between macroscopic bodies, their effect would be weak-
ened for relativistic neutrinos.
In summary, the SN 1987A signal does not seem to carry any simple
information concerning putative nongravitational long-range forces.
'x < 19 −1 −1
∼ 10 erg g s . (13.8)
Fig. 13.3. Average values for (ρ/ρ0 )n with the nuclear density ρ0 =
3×1014 g cm−3 and of (T /30 MeV)n for the protoneutron star model of
Fig. 13.2.
mostly with a time later than 0.5−1 s where the outer core has set-
tled and the shock has begun to escape. The density of the protoneu-
tron star falls within a thin shell from supranuclear levels to nearly
zero, causing the “photosphere” radius rx of the new particles to be
essentially the radius R ≈ 10 km of the settled compact star. With a
“photosphere” temperature Tx of the new objects their luminosity is
4πr2 σTx4 with the Stefan-Boltzmann constant σ which is gπ 2 /120 in
natural units with g the effective number of degrees of freedom (2 for
photons). Therefore, one must demand that
Tx <
∼ 8 MeV g
−1/4
, (13.9)
in order to stay below the total neutrino luminosity of 3×1052 erg s−1 .
It is nontrivial, however, to determine the temperature Tx which
corresponds to about unit optical depth. Following the approach of
Turner (1988) who carried this analysis through for axions one may
assume a simple model for the run of temperature and density above
the settled inner core. A simple power-law ansatz is ρ(r) = ρR (R/r)n
with the density ρR = 1014 g cm−3 at a radius R ≈ 10 km. A plausible
ansatz for the temperature profile is T (r) = TR [ρ(r)/ρR ]1/3 with TR
(temperature at radius R) of around 10 MeV. From the opacity κ as a
function of density %and temperature one may then calculate the optical
depth as τ (rx ) = r∞ x
κρ dr. From the condition τ (rx ) ≈ 23 one can
determine the “photosphere” radius rx and thus its temperature Tx .
The opacity of axions for a medium of nondegenerate nucleons was
given in Eq. (4.28). One may define τR ≡ κR ρR R so that κρR =
τR (ρ/ρR )2 (TR /T )1/2 where Eq. (4.28) yields τR = ga2 3.4×1016 . Then
one finds for Turner’s model an optical depth τa at the axion-sphere
temperature Ta
τa = τR ( 11
6
n − 1) (Ta /TR )11/2−3/n . (13.10)
Because n is a relatively large number such as 3−7 the criterion τa <
∼3
2
region where their energy flux is set. Put another way, for fermions the
concept of blackbody emission from a neutrino sphere is not adequate,
making it impossible to apply the Stefan-Boltzmann in a simplistic way.
The transport of r.h. neutrinos in the trapping limit is an equally com-
plicated problem as that of l.h. ones! Therefore, a proper treatment of
the trapping limit is generally a tricky subject; axions are the only case
where it has been studied in some detail.
Occasionally one may wish to construct a particle-physics model
that avoids the SN limit. It would be incorrect to believe that this
is achieved when the interaction strength has been tuned such that
the mean free path is of order the neutron star radius. On the con-
trary, when this condition obtains the impact on the cooling rate is
maximized. This is analogous to the impact of novel particles on the
structure and evolution of the Sun as depicted in Fig. 1.2; the cooling
rate is maximized when the mfp corresponds to a typical geometric di-
mension of the object. In the trapping regime a new particle is harmless
only if it interacts about as strongly as the particles which provide the
standard mode of energy transfer.
13.5 Axions
13.5.1 Numerical Studies
The most-studied application of the SN cooling-time argument is that
of invisible axions as these particles are well motivated (Chapter 14).
Moreover, they have attracted much interest because they are one of
the few particle-physics motivated candidates for the cosmic dark mat-
ter. Early analytic studies in the free-streaming limit are Ellis and Olive
(1987), Raffelt and Seckel (1988), and Turner (1988) who also discussed
the trapping regime; his line of reasoning was presented in Sect. 13.4.3
above. Numerical studies in the free-streaming limit were performed
by Mayle et al. (1988, 1989) and by Burrows, Turner, and Brinkmann
(1989) while the trapping regime was numerically studied by Burrows,
Ressell, and Turner (1990). The numerical studies by different work-
ers in the free-streaming limit used different assumptions concerning
the axion couplings, emission rates, and other aspects. In my pre-
vious review (Raffelt 1990d) I have attempted to reduce the results of
these works to a common and consistent set of assumptions; apart from
relatively minor differences which could be blamed on different input
physics (e.g. softer equation of state and thus higher temperatures in
the Mayle et al. papers) the results seemed reasonably consistent. A
What Have We Learned from SN 1987A? 509
recent numerical study by Keil (1994) who used the same axion emis-
sion rates as Burrows, Turner, and Brinkmann (1988) confirmed their
results.
Here, I present the numerical studies of Burrows and his collabora-
tors where axions were assumed to couple with equal strength to pro-
tons and neutrons. The axial-vector coupling to nucleons is written in
the form (C/2fa ) ψγµ γ5 ψ ∂ µ a with a model-dependent numerical factor
C, the Peccei-Quinn energy scale fa , the nucleon Dirac field ψ, and the
axion field a. Under certain assumptions detailed in Sect. 14.2.3 it can
be written in the pseudoscalar form −i ga ψγ5 ψ where ga = CmN /fa
is a dimensionless Yukawa coupling (nucleon mass mN ); Burrows et al.
used C = 12 . All results will be discussed in terms of ga and as such they
apply to any pseudoscalar particle which couples to nucleons accord-
ingly. In Sect. 14.4 the available constraints on axions will be expressed
in terms of the axion mass ma .
In the free-streaming limit the energy loss by axions was imple-
mented according to the numerical rates of Brinkmann and Turner
(1988); limiting cases of these rates were discussed in Sect. 4.2. In the
trapping regime, the transfer of energy by axions as well as axion cool-
ing from an “axion sphere” was implemented by means of an effective
radiative opacity as discussed in Sect. 4.4. The protoneutron star mod-
els are those of Burrows and Lattimer (1986) and of Burrows (1988b).
In the latter study, cooling sequences were presented for different equa-
tions of state (EOS), and different assumptions concerning the mass
and early accretion rate of the stars. A fiducial case in these studies is
model 55 with a “stiff EOS,” an initial baryon mass of 1.3 M% , and an
initial accretion of 0.2 M% .
The compatibility of a given model with the SN 1987A observations
should be tested by a maximum-likelihood analysis of the time and
energy distributions of the events in both the IMB and Kamiokande II
detectors. In practice, it is easier to consider a few simple observables.
Burrows and his collaborators chose the total number of events NKII
and NIMB in the two detectors as well as the signal duration defined by
the expected times tKII and tIMB it takes to accrue 90% of the expected
total number of events. As both detectors measured approximately
10 events each, the time of the last event probably is a reasonable
estimate of tKII and tIMB . Finally, Burrows et al. calculated the total
energy carried away by neutrinos and axions.
The run of these quantities with ga is shown in Fig. 13.4. Re-
call from Sect. 11.3.2 that the observed SN 1987A numbers of events
are NIMB = 8 and NKII = 10−12, depending on whether event No. 6
510 Chapter 13
Fig. 13.4. Results from protoneutron star cooling sequences with axions.
The free-streaming regime (small ga ) is according to Burrows, Turner, and
Brinkmann (1989), the trapping regime (large ga ) according to Burrows,
Ressell, and Turner (1990). For models A, B, and C (corresponding to
models 57, 55, and 62 of Burrows 1988b) the amount of early accretion and
the type of EOS (“stiff” or “soft”) is indicated. The models were calculated
until 20 s after collapse.
What Have We Learned from SN 1987A? 511
rate Γσ which, in the nondegenerate limit, was given in Eq. (4.7) on the
basis of a perturbative one-pion exchange (OPE) calculation. For the
protoneutron star model displayed in Fig. 13.2 the profile of this Γσ /T
is shown in Fig. 13.5. In Fig. 4.8 the axion emission rate was shown as a
function of Γσ , revealing that for the conditions of interest one is in the
neighborhood of the maximum of the solid curve. In a realistic nuclear
medium, the true spin fluctuation rate may be smaller than the OPE
calculated value, taking one perhaps somewhat to the left of the maxi-
mum. Therefore, the true axion emission rate corresponds to the naive
one (dashed line in Fig. 4.8) at Γσ /T ≈ 3−5 which at temperatures
around 30 MeV corresponds to around 20% nuclear density.
Fig. 13.5. Profile for the nondegenerate spin-fluctuation rate Γσ of Eq. (4.7)
in the protoneutron star model S2BH 0 of Keil, Janka, and Raffelt (1995)
shown in Fig. 13.2.
Fig. 13.6. Number of events NKII and NIMB in the Kamiokande and IMB
detectors as well as the signal duration tKII and tIMB (in sec) as a function
of the assumed number of neutrino flavors (Burrows, Ressell, and Turner
1990). The signal duration is defined as the time it takes to accrue 90% of
the total expected number of events.
514 Chapter 13
Fig. 13.7. Number of events in the Kamiokande and IMB detectors as well
as the signal durations as a function of the assumed “opacity suppression
parameter” defined by Eq. (13.12). Filled circles refer to a suppression of
both neutral- and charged-current axial-vector interactions while open circles
refer to a suppression of neutral-current interactions only. (Adapted from
Keil, Janka, and Raffelt 1995.)
too different from that found in the dilute-medium limit. If one applies
the analytic criterion Eq. (13.8) one finds
mν <
∼ 30 keV (13.15)
where mWR,L are the r.h. and l.h. charged gauge boson masses while ζ
is the left-right mixing parameter.
In order to constrain 'CC one assumes the existence of r.h. νe ’s so
that the dominant energy-loss mechanism of a SN core is e + p →
n + νe,R where the final-state r.h. neutrino escapes freely. Initially, a
substantial fraction of the thermal energy of a SN core is stored in
the degenerate electron sea. Therefore, the time scale of cooling is
estimated by the inverse scattering rate for e + p → n + νe,R . The usual
charged-current weak scattering cross section involving nonrelativistic
nucleons is G2F (CV2 +3CA2 )Ee2 /π with CV2 +3CA2 ≈ 4 in a nuclear medium
(Appendix B). Using a proton density corresponding to nuclear matter
at 1015 g cm−3 and using 100 MeV for a typical electron energy one finds
520 Chapter 13
(e.g. Jodidio et al. 1986) which is much weaker but does not depend
on the assumed existence of r.h. neutrinos. Mohapatra and Nussinov
(1989) extended the SN 1987A bound to the case of r.h. Majorana
neutrinos which mix with νe .
In order to constrain r.h. neutral currents, equivalent to constraining
the mass of putative r.h. Z ◦ gauge bosons, one considers the emission
of r.h. neutrino pairs νR ν R . The dominant emission process is by the
nucleons of the medium; in a dilute medium it can be represented as the
bremsstrahlung process N N → N N νR ν R . Apart from a global scaling
factor '2NC , the bremsstrahlung energy-loss rate for a nondegenerate
medium was given in Eq. (4.23). However, in a dense medium this
rate probably saturates at around 10% nuclear density as in the case
of axion emission (Sect. 4.6.7). Evaluating Eq. (4.23) at 10% nuclear
density (ρ15 = 0.03) and at T = 30 MeV, and applying the analytic
criterion Eq. (13.8) one finds
'NC < −3
∼ 3×10 . (13.19)
This is less restrictive from what was found by Raffelt and Seckel (1988)
or Barbieri and Mohapatra (1989).
The translation of a limit on 'NC into one on a r.h. gauge boson mass
depends on details of the couplings to quarks and leptons, and notably
on the mixing angle between the new and the standard Z bosons. De-
tailed analyses were presented by Grifols and Massó (1990b), Grifols,
Massó, and Rizzo (1990), and Rizzo (1991). Because these authors did
not consider multiple-scattering effects and the resulting saturation of
the bremsstrahlung process, their bounds on the Z $ mass are somewhat
too restrictive, perhaps by a factor of 2 or 3. Still, mZ ! has to exceed
at least 1 TeV, except for special choices of the mixing angle.
The SN 1987A limits on r.h. neutral currents are weaker than those
from big bang nucleosynthesis ('NC < −3
∼ 10 ) which are based on the
requirement that r.h. neutrinos must not have come to thermal equi-
librium after the QCD phase transition (at T < ∼ 200 MeV) in the early
universe (e.g. Olive, Schramm, and Steigman 1981; Ellis et al. 1986).
What Have We Learned from SN 1987A? 521
µν < −12
∼ 4×10 µB (13.20)
As the r.h. neutrinos escape from the inner core with much larger ener-
gies than those from the neutrino sphere one would expect high-energy
events in the Kamiokande and IMB detectors, contrary to the observa-
tions. Therefore, one probably needs to require µν < −12
∼ 10 µB for the
diagonal dipole moments; spin-flavor oscillations could be suppressed
by the neutrino mass differences. As the spin precession is the same for
all neutrino energies, this limit would not apply if the Earth happened
to be in a node of the oscillation pattern between SN 1987A and us.
Neutrino magnetic moments of order 10−12 µB could also affect the
infall phase of SNe. The spin-flip scattering on nuclei would be coher-
ently enhanced relative to protons. Therefore, neutrinos could escape
in the r.h. channel for much longer so that effectively trapping would
set in much later than in the standard picture (Nötzold 1988).
In and near the SN core there probably exist strong magnetic fields
of order 1012 Gauss or more which would induce spin-precessions be-
tween r.h. and l.h. neutrinos. Therefore, the sterile states produced in
the deep interior by spin-flip scattering could back-convert into active
ones near the neutrino sphere. Depending on details of the matter-
induced neutrino energy shifts, the vacuum mass differences, and the
magnetic field strengths and configurations this conversion could take
place inside or outside of the neutrino sphere. The observable neutrino
signal could be affected, but also the energy transfer within the SN core
and outside of the neutrino sphere. Perhaps, a more efficient transfer of
energy to the stalled shock wave could help to explode SNe in the de-
layed explosion scenario. Various aspects of these scenarios have been
studied by Dar (1987), Nussinov and Rephaeli (1987), Goldman et al.
(1988), Voloshin (1988), Okun (1988), Blinnikov and Okun (1988), and
Athar, Peltoniemi, and Smirnov (1995).
Clearly, Dirac magnetic or transition moments in the 10−12 µB range
and below would affect SN dynamics and the observable neutrino signal
in interesting ways. However, because there are so many parameters
and possible field configurations, it is hard to develop a clear view of
the excluded or desired neutrino properties. If compelling evidence for
nonstandard neutrino electromagnetic properties in this range were to
emerge, SN dynamics likely would have to be rethought from scratch.
13.8.4 Millicharges
Within the particle physics standard model it is not entirely impos-
sible that neutrinos have small electric charges (Sect. 15.8). In this
case neutrinos would have to be Dirac fermions and so the r.h. states
What Have We Learned from SN 1987A? 523
eν < −9
∼ 10 e. (13.21)
Axions
524
Axions 525
Fig. 14.1. Axion mixing with qq states and thus π ◦ . The curly lines represent
gluons, the solid lines quarks.
The ground state of the axion field is at the minimum of its po-
tential at a = 0, explaining the absence of a neutron electric dipole
moment. If one could produce a static nonvanishing axion field a0 in
some region of space, neutrons there would exhibit an electric dipole
moment corresponding to Θ = −a0 /fa .
The Lagrangian Eq. (14.3) is the minimal ingredient for any ax-
ion model: the aGG ! coupling is their defining feature as opposed to
other pseudoscalar particles. Then axions inevitably acquire an effec-
tive mass at low energies. Thus the concept of a “massless axion” for
some arbitrary pseudoscalar is a contradiction in terms.
528 Chapter 14
Because of the mixing with π ◦ , axions share not only their mass, but
also their couplings to photons and nucleons with a strength reduced
by about fπ /fa . Therefore, they generically couple to photons so that
the general discussion of Chapter 5 applies directly except for those
aspects which required massless pseudoscalars.
The effective axion mass is a low-energy phenomenon below ΛQCD ≈
200 MeV. Above this energy pions and other hadrons dissociate in
favor of a quark-gluon plasma. Then a = 0 is no longer singled out so
that any value in the interval 0 ≤ a < 2πfa is physically equivalent.
Because the universe is believed to begin with a hot and dense “big
bang,” any initial value for a is equally plausible, or different initial
conditions in different regions of space. As the universe expands and
cools below ΛQCD , however, the axion field must relax to its newly
singled-out ground state at a = 0. This relaxation process produces a
population of cosmic background axions which is, in units of the cosmic
critical density (e.g. Kolb and Turner 1990),
Ωa h2 ≈ (fa /1012 GeV)1.175 . (14.5)
The exact value depends on details of the cosmic scenario and of the
relaxation process. Modulo this uncertainty, values exceeding fa ≈
1012 GeV are excluded as axions would overdominate the dynamics of
the universe. With Eq. (14.4) this corresponds to ma < −5
∼ 10 eV; axions
near this bound would be the cosmic dark matter. A search strategy
for galactic axions in this mass range was discussed in Sect. 5.3.
of a with gluons is then given by the triangle graph of Fig. 14.2. With
the first term of Eq. (14.10) it yields an effective a-gluon interaction of
ga α s !,
LaG = − a GG (14.11)
m 8π
where αs ≡ gs2 /4π. All external momenta were taken to be small relative
to the mass m of the loop fermion.
Fig. 14.2. Triangle loop diagram for the interaction of axions with gluons
(strong coupling constant gs , axion-fermion Yukawa coupling ga ). An anal-
ogous graph pertains to the coupling of axions with photons if the fermion
carries an electric charge which replaces gs .
The total aGG ! interaction is obtained as a sum over Eq. (14.11) for all
Ψj . Because gaj = Xj mj /fPQ the fermion masses drop out. With
%
N≡ j
Xj and fa ≡ fPQ /N (14.13)
one has then found the required coupling Eq. (14.3) which allows one
to interpret a as the axion field.
The potential V (a) is periodic with 2πfa = 2πfPQ /N . The interpre-
tation of a as the phase of Φ, on the other hand, implies a periodicity
with 2πfPQ so that N must be a nonzero integer. This requirement
restricts the possible assignment of PQ charges to the quark fields. It
also implies that there remain N different equivalent ground states for
the axion field, each of which satisfies Θ = 0 and thus solves the CP
problem.
Axions 531
with the axion example Eq. (14.15) where only one Nambu-Goldstone
boson was present as opposed to the pion isotriplet. However, this ad-
ditional term does not contribute to the bremsstrahlung process in the
limit of nonrelativistic nucleons so that the above conclusion regarding
the derivative coupling remains valid.
In the process N N → N N a (Fig. 14.3) axions and pions appear so
that again two Nambu-Goldstone bosons are attached to one fermion
line. It is then necessary to use a derivative coupling for at least one
of them (Raffelt and Seckel 1988). For other bremsstrahlung processes
such as e− p → pe− a, where the particles interact through a virtual
photon (a gauge boson) the pseudoscalar coupling causes no trouble.
Also for the Compton process γe− → e− a one may use either the pseu-
doscalar or the derivative axion coupling: both yield the same result.
Because it is not always a priori obvious whether the pseudoscalar and
derivative couplings yield the same result it is a safe strategy to use the
derivative coupling in all calculations.
Notably, the Higgs field Φ which gives rise to the axion probably
exhibits effective interactions of dimension 2m + n
(ΦΦ† )m Φn
Vgrav (Φ) = g eiδ , (14.17)
m2m+n−4
Pl
where g and δ are real numbers. Because such interactions violate the
PQ symmetry for n += 0 they induce an effective potential for the axion
after spontaneous symmetry breaking. The full potential is then of the
form (Kamionkowski and March-Russell 1992)
V (a) & ' & '
2 2
= m QCD 1 − cos(a/f a ) + m grav 1 − cos(δ + na/f a ) ,(14.18)
fa2
where mQCD is the usual QCD axion mass while gravity induces
√
m2grav = g m2Pl (fa / 2 mPl )2m+n−2 . (14.19)
For g of order unity and for low values of m and n one needs a very
small fa for the QCD effect to dominate. Therefore, unless gravity for
some reason favors a minimum at the CP-conserving position for a the
PQ scheme will be ruined entirely.91
Whatever the ultimate quantum theory of gravitation, no doubt it
will be very special. Therefore, it is by no means obvious that the
above arguments, which do not go far beyond a dimensional analysis,
correctly represent the low-energy effects of Planck-scale physics. Even
then the PQ mechanism still works if the PQ global symmetry is an
“automatic symmetry” of a gauge theory; in this case it is protected
from the assault of quantum gravity. Such models can be constructed
(Holman et al. 1992) and in fact may be quite generic (Barr 1994).
Either way, in order for axions to solve the strong CP problem one
must assume that the PQ scheme is not ruined by quantum gravity.
This discussion illustrates an important feature of axion models,
or any model involving a broken global symmetry and its Nambu-
Goldstone boson. These particles are interlopers in the low-energy
world—axions really belong to the high-energy world at the PQ scale.
These roots make them susceptible to physics at large energy scales,
at the Planck mass, for example. By the same token, if axions were
ever detected, for example by the galactic axion search (Sect. 5.3), they
would be one of the few messengers that we can ever hope to receive
from a high-energy world which is otherwise inaccessible to experimen-
tal enquiry.
91
These issues were studied by Barr and Seckel (1992) and by Kamionkowski and
March-Russell (1992). See also the earlier papers by Georgi, Hall, and Wise (1981),
Lazarides, Panagiotakopoulos, and Shafi (1986), and Dine and Seiberg (1986).
534 Chapter 14
Axions generically mix with pions so that their mass and their couplings
to photons and nucleons are crudely fπ /fa times those of π ◦ . In detail,
however, these properties depend on the specific implementation of
the PQ mechanism. Therefore, it is useful to review briefly the most
common axion models which may serve as generic examples for an
interpretation of the astrophysical evidence.
In the standard model, the would-be Nambu-Goldstone boson from
the spontaneous breakdown of SU(2)×U(1) is interpreted as the third
component of the neutral gauge boson Z ◦ , making it impossible for the
scalar field Φ of which axions are the phase to be the standard Higgs
field. Therefore, one needs to introduce two independent
√ √Higgs fields Φ1
and Φ2 with vacuum expectation √ values f 1 / 2 and f 2 / 2 which must
2 2 1/2 −1/2
obey (f1 + f2 ) = fweak ≡ ( 2 GF ) ≈ 250 GeV. In this standard
axion model (Peccei and Quinn 1977a,b; Weinberg 1978; Wilczek 1978)
Φ1 gives masses to the up- and Φ2 to the down-quarks and charged
leptons. With x ≡ f1 /f2 and 3 families the axion decay constant is
fa = fweak [3 (x + 1/x)]−1 <
∼ 42 GeV. This and related “variant” models
(Peccei, Wu, and Yanagida 1986; Krauss and Wilczek 1986), however,
are ruled out by overwhelming experimental and astrophysical evidence;
for reviews see Kim (1987), Cheng (1988), and Peccei (1989).
Therefore, one is led to introduce an√electroweak singlet Higgs field
with a vacuum expectation value fPQ / 2 which is not related to the
weak scale. Taking fPQ , fweak , the mass of the axion becomes
very small, its interactions very weak. Such models are generically re-
ferred to as invisible axion models. The first of its kind was the KSVZ
model (Kim 1979; Shifman, Vainshtein, and Zakharov 1980) discussed
in Sect. 14.2.2. It is very simple because the PQ mechanism entirely
decouples from the ordinary particles: at low energies, axions interact
with matter and radiation only by virtue of their two-gluon coupling
which is generic for the PQ scheme. The KSVZ model in its simplest
form is determined by only one free parameter, fa = fPQ , although one
may introduce N > 1 exotic quarks whence fa = fPQ /N .
Also widely discussed is the DFSZ model introduced by Zhitnitskiı̆
(1980) and by Dine, Fischler, and Srednicki (1981). It is a hybrid be-
tween the standard and KSVZ models in that it uses an √ electroweak
singlet scalar field Φ with a vacuum expectation value fPQ / 2 and two
electroweak doublet fields Φ1 and Φ2 . There is no need, however, for ex-
Axions 535
otic heavy quarks: only the known fermions carry Peccei-Quinn charges.
Therefore, N is the number of standard families. Probably N = 3 so
that the remaining free parameters of this model are fa = fPQ /N and
x = f1 /f2 which is often parametrized by x = cot β or equivalently by
cos2 β = x2 /(x2 + 1).
From a practical perspective, the main difference between the KSVZ
and DFSZ models is that in the latter axions couple to charged leptons
in addition to nucleons and photons. The former is an example for the
category of hadronic axion models.
Because fPQ , fweak in these models, one may attempt to identify
fPQ with the grand unification scale fGUT ≈ 1016 GeV (Wise, Georgi,
and Glashow 1981; Nilles and Raby 1982). However, the cosmologi-
cal bound fa < 12
∼ 10 GeV disfavors the GUT assignment. There exist
numerous other axion models, and many attempts to connect the PQ
scale with other scales—for a review see Kim (1987). In the absence
of a compelling model fa should be viewed as a free phenomenological
parameter.
The axion mass which arises from its mixing with π ◦ can be obtained
with the methods of current algebra to be (Bardeen and Tye 1978;
Kandaswamy, Salomonson, and Schechter 1978; Srednicki 1985; Georgi,
Kaplan, and Randall 1986; Peccei, Bardeen, and Yanagida 1987)
( )1/2
fπ mπ z
ma =
fa (1 + z + w)(1 + z)
107 GeV
= 0.60 eV , (14.20)
fa
where the quark mass ratios are (Gasser and Leutwyler 1982)
Axions which interact too strongly to escape freely from the interior
of stars would still contribute to the transfer of energy. For the Sun,
this issue was studied in Sect. 1.3.5. One easily finds that for Ce = 1
Fig. 1.2 excludes axion masses below about 50 keV.
In hadronic axion models Ce = 0 at tree level and so no interesting
bounds on ma and fa obtain. In the DFSZ model, Ce was given in
Eq. (14.28). Taking the number of families to be Nf = 3 one finds
ma ξ < > 7
∼ 0.4 eV and fa /ξ ∼ 1.5×10 GeV, (14.35)
emitted from an “axion sphere” rather than the entire volume of the
protoneutron star.
These bounds were derived assuming equal couplings to protons and
neutrons. However, a glance at Fig. 14.4 reveals that KSVZ axions es-
sentially do not couple to neutrons while Cp ≈ −0.36. For DFSZ axions
the couplings vary with cos2 β, although for cos2 β ≈ 0.5 about the same
values as for KSVZ axions apply which are thus taken as generic. As-
suming a proton fraction of about 0.3 for the relevant regions of the SN
core I estimate an effective nucleon coupling of CN ≈ 0.31/2 0.36 ≈ 0.2.
Therefore,
0.01 eV < <
∼ ma ∼ 10 eV,
0.6×106 GeV < < 9
∼ fa ∼ 0.6×10 GeV (14.37)
are formally adopted as the SN 1987A excluded axion parameters.
Axions on the “trapping side” of the SN argument can still be ex-
cluded because they would have caused additional events in the IMB
and Kamiokande water Cherenkov detectors. The excluded range of
Eq. (13.2) translates into the approximate mass exclusion range of
20 eV−20 keV.
These limits on the axion mass and decay constant are summarized
in Fig. 14.5. The slanted end of the SN 1987A exclusion bar is a
reminder of the potentially large uncertainty of this limit. Except for
very special choices of model-dependent parameters axions with a mass
above 0.01 eV are excluded.
The high-mass end of the stellar exclusion bars in Fig. 14.5 has
not been worked out in detail because of their overlap with laboratory
limits. The globular cluster limits apply without modification up to a
mass of, say, 30 keV because the temperature in the cores of HB stars
and red giants are about 10 keV. However, axions with masses in this
range interact much more strongly than those at the low-mass end of the
exclusion bar so that the Boltzmann suppression of the emission rate
for a large mass is partly balanced by the increased coupling strength.
Miscellaneous Exotica
545
546 Chapter 15
Table 15.1. Bounds on the present-day ĠN /GN . (Adapted from Will 1993.)
Table 15.2. Characteristics of the Demarque et al. (1994) solar models with
a varying GN according to Eq. (15.1).
37
β α Xinitial Xc Tc ρc Renv Cl 71 Ga
[%] [%] [106 K] [g/cm3 ] [R" ] [SNU] [SNU]
−0.4 1.832 69.42 45.6 15.14 125.5 0.739
−0.2 1.895 70.10 42.1 15.28 134.0 0.731 5.5 117
−0.1 1.936 70.51 40.0 15.37 139.5 0.724
0.0 1.983 70.98 37.6 15.47 146.2 0.721 6.8 124
0.1 2.036 71.51 34.8 15.58 154.6 0.716
0.2 2.104 72.11 31.7 15.72 165.1 0.710 8.7 134
0.4 2.291 73.56 23.9 16.07 197.3 0.695
where Γ0 = ĠN (t0 )/GN (t0 ) is the present-day rate of change of Newton’s
constant. Then one finds explicitly
τ γ 1 Γ0 τ 1 − (1 − γ1 Γ0 τ∗ )1/γ1
= = , (15.5)
τ∗ 1 − (1 − Γ0 τ )γ1 Γ0 τ∗
where γ1 ≡ γ + 1. Given a present-day rate of change Γ0 one can
thus determine the modification of the globular-cluster age if a certain
apparent age or a certain true age is assumed.
The observed color-magnitude diagrams of globular clusters yield
apparent ages τ∗ in the range 14 to 18 Gyr. With these values one
Miscellaneous Exotica 553
Fig. 15.1. Required present-day ĠN /GN in order to achieve a true globular-
cluster age τ , given that the apparent age is τ∗ . A linear GN (t) variation as
in Eq. (15.4) was assumed.
can relate a desired true age τ to a required value for Γ0 (Fig. 15.1).
Conversely, it is probably safe to assume that the true ages of globular
clusters do not exceed 20 Gyr. Then Fig. 15.1 implies that today
A lower age limit is less certain. Taking 8 Gyr one finds ĠN /GN > ∼
−35×10−12 yr−1 which is less certain, and also less interesting relative
to the limits discussed in the previous sections.
Fig. 15.2. Summary of limits on the present-day ĠN /GN . The big-bang
nucleosynthesis limit is not shown as it depends sensitively on the assumed
GN (t) variation at early cosmic times.
554 Chapter 15
(Bouquet and Vayonakis 1982; Fukugita and Sakai 1982; Anand et al.
1984). The neutrino burst of SN 1987A yielded more interesting limits
for the case of low-mass photinos (Ellis et al. 1988; Grifols, Massó, and
Peris 1989; Grifols and Massó 1990b). To avoid that too much energy
is carried away by photinos they inferred that squark masses in the
approximate range 60 GeV to 2.5 TeV were excluded.
Low-mass neutralinos are disfavored by laboratory limits. More-
over, if the LSP plays the role of cold dark matter its mass likely is
above several 10 GeV. In this case the stellar energy-loss arguments
would not yield any constraints as all supersymmetric particles would
be too heavy to be emitted. Stars would still play an interesting role as
they could trap the dark-matter particles. Their annihilation in the Sun
or Earth would lead to a high-energy neutrino signal which has been
constrained by the Kamiokande detector (Mori et al. 1992). It may
well be found at the Cherenkov detectors Superkamiokande, NESTOR,
DUMAND, or AMANDA and thus lead to the indirect discovery of par-
ticle dark matter in the galaxy. These important issues are discussed
at length in the forthcoming review Supersymmetric Dark Matter by
Jungman, Kamionkowski, and Griest (1995).
15.7 Majorons
15.7.1 Particle-Physics and Cosmological Motivations
Axions (Chapter 14) are one representative of a variety of Nambu-
Goldstone bosons of spontaneously broken global symmetries that have
appeared in the literature over the years. Another widely discussed ex-
ample are the majorons first introduced by Chicashige, Mohapatra,
and Peccei (1981) as a scheme to generate small neutrino Majorana
masses. An important variation by Gelmini and Roncadelli (1981) and
Georgi, Glashow, and Nussinov (1981) led to a model where neutri-
nos had small Majorana masses and coupled to the massless majoron
(a pseudoscalar boson like the axion) with a relatively large Yukawa
strength. The main phenomenological interest in this sort of conjec-
ture lies in the intriguing possibility that neutrinos could have relatively
strong interactions with the majorons and with each other by virtue of
majoron exchange. As majorons would not necessarily show up in in-
teractions with ordinary matter one could well speculate that neutrinos
might have “secret interactions” which would be of relevance only in
a neutrino-dominated environment such as the early universe, perhaps
the present-day universe if neutrinos have a cosmologically significant
Miscellaneous Exotica 559
masses by a see-saw type mixing effect with the heavy states. The ma-
joron coupling to standard neutrinos would be extremely small in this
model, leading to no interesting consequences besides small Majorana
masses for νe , νµ , and ντ .
Gelmini and Roncadelli (1981) and Georgi, Glashow, and Nussinov
(1981) suggested instead to do away with the unobserved heavy sterile
neutrinos and give a small Majorana mass directly to the sequential
neutrinos by the interaction with the new Higgs field. The intriguing
feature of this model is that it requires a very small vacuum expectation
value v, perhaps in the keV regime. As all couplings of the new Higgs
field to fermions scale with the inverse of v, the majoron would have
a rather strong coupling to neutrinos. Among many fascinating phe-
nomenological and astrophysical consequences (e.g. Georgi, Glashow,
and Nussinov 1981; Gelmini, Nussinov, and Roncadelli 1982) this model
predicted, however, that the new Higgs field should contribute precisely
the equivalent of two massless neutrino species to the Z ◦ decay width.
The measurements of this width at CERN and SLAC in 1989−1990,
however, correspond exactly to the known three neutrino flavors (Parti-
cle Data Group 1994), leaving no room for this “triplet majoron model.”
Other “doublet majoron models” which would contribute one-half of an
effective neutrino species to the Z ◦ decay width are also excluded (for
references see, e.g. Berezhiani, Smirnov, and Valle 1992). It is possible,
however, to construct majoron models for Majorana neutrino masses
which retain the original idea of Chicashige, Mohapatra, and Peccei
(1981) and yet provide large majoron-neutrino couplings (e.g. Berezhi-
ani, Smirnov, and Valle 1992 and references therein; see also Burgess
and Cline 1994a; Kikuchi and Ma 1994, 1995).
The main motivation for going out of one’s way to construct such
models does not arise from particle theory but rather from experi-
ments and astrophysics. In Sect. 7.1.4 it was outlined that those nuclei
which decay predominantly by a double beta channel (emission of 2e−
and 2ν e ) can also decay in a neutrinoless mode if νe has a Majorana
mass, allowing an emitted ν e to be effectively reabsorbed as a νe . In
majoron models of Majorana neutrino masses there is a third decay
channel where the intermediate νe in the 0ν mode radiates a majoron
so that effectively 2e− plus one majoron χ are emitted. The expected
sum spectrum of the electron energies would be continuous as in the
2ν mode, but with a different spectral shape. Once in a while, ex-
periments which search for the 0ν mode (a sharp endpoint peak of
the 2e− sum spectrum) have reported a continuous spectral signature
which allegedly could not be ascribed to the dominant 2ν mode or other
Miscellaneous Exotica 561
formation and thus could be a novel ingredient for cold dark matter
cosmological models (Dodelson, Gyuk, and Turner 1994).
Even if in the long run the 2β experiments do not yield any com-
pelling evidence for majoron emission, one may consider a decaying-
neutrino cosmology as a motivation in its own right for majoron mod-
els. Such cosmologies may explain the discrepancy between the cosmic
density fluctuation spectrum inferred from the cosmic microwave back-
ground and from galaxy correlations which persists in a purely cold
dark matter cosmology (Bond and Efstathiou 1991; Dodelson, Gyuk,
and Turner 1994; White, Gelmini, and Silk 1995). Another neutrino-
related explanation of this discrepancy is a hot plus cold dark matter
cosmology which involves neutrinos with a mass of a few eV.
In summary, the simplest majoron model which implied large cou-
plings to neutrinos (the Gelmini-Roncadelli model) is experimentally
excluded although one can construct more complicated ones which re-
tain sizeable neutrino-majoron couplings and yet are compatible with
the Z ◦ decay width. The possibility of such models is entertained be-
cause 2β decay experiments may yet turn up compelling evidence for
majoron decays, and because certain cosmological models of structure
formation may be taken to suggest massive, decaying neutrinos.
the present author has not been able to develop a clear view of the
precise range of parameters that can be ruled out or ruled in by the
SN 1987A neutrino signal.97
eνµ < −9
∼ 10 e. (15.9)
where ex and mx are the charge and mass of the millicharged particle,
respectively. This result applies to mx >∼ 1 keV.
Davidson, Campbell, and Bailey (1991) have reviewed more restric-
tive bounds from a host of accelerator experiments (Fig. 15.3).
A simple astrophysical constraint is based on avoiding excessive en-
ergy losses of stars which can produce millicharged particles by various
reactions, most notably the plasma decay process. To avoid an unac-
ceptable delay of helium ignition in low-mass red giants, and to avoid an
566 Chapter 15
568
Neutrinos: The Bottom Line 569
Table 16.1. Approximate neutrino parameters which explain all current solar
neutrino observations in the framework of standard solar model assumptions.
The main aspect of the current situation is that there does not seem
to be a simple “astrophysical solution” to reconcile the solar source
spectrum with the measured fluxes in experiments with three different
spectral response characteristics. Even allowing for large modifications
of the p 7 Be cross section or the solar central temperature does not yield
consistency unless one stretches the experimental uncertainties of the
flux measurements beyond reasonable limits. Still, a final verdict on
the question of solar neutrino oscillations can be expected only from
the near-future experiments Superkamiokande, SNO, and BOREXINO
as discussed in Chapter 10.
state general constraints as one may easily postulate, for example, that
the decays do not involve final-state ν e ’s, that even “wrong-helicity”
Dirac neutrinos are trapped in a SN core by novel interactions, or that
heavy Majorana ντ ’s have only negligible mixings with νe . Surely other
loopholes could be found.
b) Laboratory Limits
Less problematic bounds on neutrino dipole moments arise from labo-
ratory experiments where one studies the recoil spectrum of electrons
in the reaction ν + e → e + ν ! where ν ! can be the same or a different
flavor (Sect. 7.5.1). A sensitivity down to, perhaps, as low as 10−11 µB
can be expected from a current effort involving reactor neutrinos as
a source (MUNU experiment). Current limits on dipole or transition
moments are about 2×10−10 µB if νe is involved, and about 7×10−10 µB
if νµ is involved. For transition moments, these limits are subject to
the assumption that there is no cancellation between a magnetic and
an electric dipole scattering amplitude.
16.3.4 Summary
Neutrinos with nonstandard interactions may well saturate the experi-
mental mass limits, and may have a variety of novel properties. How-
ever, it is nearly impossible to derive generic constraints on quanti-
ties like magnetic transition moments without specifying an underlying
particle physics model. Many constraints, notably those related to
SN 1987A or to cosmology, can be circumvented by postulating suf-
ficiently bizarre neutrino properties. Therefore, it is probably more
important to know the arguments that can serve to learn something
about neutrinos in astrophysics than it is to know a list of alleged lim-
its. If a concrete conjecture turns up, or a specific theoretical model
needs to be constrained, one can easily go through the list of arguments
and check if they apply or not. Perhaps this book can be of help at
this task.
Appendix A
580
Units and Dimensions 581
cm s ly pc
cm 1 0.334×10−10 1.06×10−18 0.325×10−18
s 2.998×1010 1 0.317×10−7 0.973×10−8
ly 0.946×1018 3.156×107 1 0.307
pc 3.08×1018 1.028×108 3.26 1
a
Atomic mass unit.
Appendix B
Neutrinos can interact with other fermions and with each other by
the exchange of W or Z bosons. Because the astrophysical phenomena
relevant for this book take place at very low energies compared with the
W or Z mass, one may always use an effective four-fermion coupling
which is parametrized in terms of the Fermi constant and the weak
mixing angle
GF = 1.166×10−5 GeV−2 ,
sin2 ΘW = 0.2325 ± 0.0008. (B.1)
583
584 Appendix B
Table B.1. Neutral-current couplings for the effective Hamiltonian Eq. (B.4)
in vacuum.
Numerical Neutrino
Energy-Loss Rates
In normal stars with densities below nuclear there are four main reac-
tions that contribute to the energy loss by neutrino emission:
γpl → νν Plasma process,
γ e− → e− νν Photoneutrino process,
(C.1)
e+ e− → νν Pair annihilation,
e− (Z, A) → (Z, A) e− νν Bremsstrahlung.
Individual processes dominate in the regions of density and temperature
indicated in Fig. C.1. In the following, various analytic fit formulae for
these neutrino emission rates are reviewed.
585
586 Appendix C
Fig. C.1. Regions of density and temperature where the indicated neutrino
emission processes contribute more than 90% of the total. µe is the electron
“mean molecular weight,” i.e. roughly the number of baryons per electron.
The bremsstrahlung contribution depends on the chemical composition. The
solid lines are for helium, the dotted ones for iron which yields a larger
bremsstrahlung rate.
Fig. C.2. Deviation between the Schinder et al. (1987) and the Itoh et al.
(1989) rates for the photoneutrino and pair-annihilation processes. Com-
pared are the total emission rates where the plasma rate of Haft, Raffelt and
Weiss (1994) and the bremsstrahlung rate (helium) of Itoh and Kohyama
(1983) were used. The contours indicate were the individual processes dom-
inate (Fig. C.1).
energy-loss rates for the photo and pair process was calculated accord-
ing to these authors, while in each case the plasma rate of Haft, Raffelt,
and Weiss (1994) and the bremsstrahlung rate for helium of Itoh and
Kohyama (1983) were taken. Therefore, deviations occur only in the
range of temperatures and densities where the photo or pair process
dominates. The largest deviations in the lower left corner of Fig. C.2
are around 25%. However, there the absolute magnitude of neutrino
emission is very small (see below) so that the difference between the
rates in this regime does not appear to be of much practical significance.
Another analytic approximation formula for the pair process was
derived by Blinnikov and Rudzskiı̆ (1989).
588 Appendix C
C.3 Bremsstrahlung
Bremsstrahlung dominates for low temperatures and high densities
where electrons are degenerate and the nuclei are strongly correlated.
In a series of papers the emission rate was calculated by Itoh and Ko-
hyama (1983), Itoh et al. (1984a,b), and Munakata, Kohyama, and Itoh
(1987).
For simple estimates one may use the approximate rate given in
Eq. (11.40). In Fig. C.3 I display the error of this approximation for
iron relative to the results of Itoh and Kohyama (1983). For orientation
the contours of Fig. C.1 for iron are also shown, but only the brems-
strahlung rates are compared. The simple approximation is not a bad
fit in the regions where bremsstrahlung could be of interest. For carbon
the fit is almost as good, but it is substantially worse for helium.
Fig. C.3. Relative deviation between the bremsstrahlung rates of Itoh and
Kohyama (1983) for iron and the simple approximation formula Eq. (11.40).
The contours where different processes dominate are for iron.
Numerical Neutrino Energy-Loss Rates 589
Fig. C.4. Contour plot for the total neutrino energy-loss rate !ν per unit
mass for helium. The thin contours are at intervals of a factor of 10 for !ν .
The regions where the individual processes dominate are also indicated.
590 Appendix C
Fig. C.5. Neutrino energy-loss rate as a function of density. The thin lines are
for temperatures 2, 3, 4, etc. times the value indicated on the corresponding
thick line.
Characteristics of Stellar
Plasmas
591
592 Appendix D
Ye ≈ µ−1 1
e ≈ X + 2 (Y + Z), (D.3)
where Ye is the mean number of electrons per baryon. Here, the mass
fraction Z of metals must not be confused with a nuclear charge.
Some examples for typical conditions encountered in stars are shown
in Fig. D.1, ignoring neutron stars (density around nuclear). Aside
from the example of an evolved massive star, all conditions refer to the
centers of stars. Except for the hydrogen main squence, the abscissa is
essentially the physical density because for most chemical compositions
Ye ≈ 12 . Horizontal-branch (HB) stars correspond essentially to the
helium main sequence at 0.5 M" .
Fig. D.2. Contours for $v 2 %1/2 , the average thermal velocity of electrons.
The loci of the stellar models of Fig. D.1 are also indicated.
Fig. D.3. Contours for the electron chemical potential µ. The solid lines are
marked with the relevant value for µ, the dotted lines with µ − me .
For only one species of ions with charge Ze one has ki2 = Z kD
2
. Nu-
merically,
Γ = Z 2 α/ai T, (D.16)
with the mass density in units of g cm−3 and T8 = T /108 K. Recall that
for a single nuclear species Ye = Z/A (atomic weight A).
The plasma is weakly coupled for Γ < ∼ 1, it is in the liquid metal
<
phase for 1 ∼ Γ < 178, and forms a body-centered cubic lattice for
Γ > 178 (Slattery, Doolen, and DeWitt 1980, 1982). Debye screening
by the ions is appropriate for a weakly coupled plasma; otherwise the
Debye approximation for the ion-ion correlations is misleading. Con-
tours for Γ in the ρ-T -plane are shown in Fig. D.6. For the purposes
of this book, a strongly coupled plasma occurs only in the interior of
white dwarfs.
Fig. D.6. Contours for the plasma coupling parameter Γ. Also shown are
the loci of the stellar models of Fig. D.1 which had to be shifted relative
to each other according to the nuclear charge Z relevant for each chemical
composition.
D.1.6 Summary
The characteristic plasma properties for a number of typical astrophys-
ical sites that are important in the main body of the book are summa-
rized in Tab. D.1.
Table D.1. Plasma characteristics for some typical astrophysical sites.
nB = nn + np Baryon density,
np = ne Charge neutrality, (D.18)
µe + µp = µn + µνe β equilibrium.
x4n
x2p = . (D.24)
4 (1 + x2n )
This result may be expressed in terms of the usual composition param-
eters Yp and Yn which give the number of protons and neutrons per
baryon; Yp + Yn = 1. Then Yn,p = (xn,p /xB )3 so that
# $3
xB (1 − Yp )2
Yp = . (D.25)
2 [1 + x2B (1 − Yp )2/3 ]3/2
Fig. D.7. (a) Contours for Yp in hot neutron-star matter with YL = 0.3.
Solid lines for the effective nucleon mass as in Fig. D.7, dotted lines for the
vacuum mass.
Characteristics of Stellar Plasmas 603
Fig. D.9. Ratio of the suppression factor of Fig. D.8 between the case with
an effective nucleon mass and with the vacuum one.
YL = 0.3 and the effective nucleon mass of Fig. 4.10. In Fig. D.9 the
ratio of this factor between the case of an effective nucleon mass and
the vacuum one are shown, i.e. the additional Pauli suppression of the
neutrino scattering rate from using an effective nucleon mass.
References
Prefixes to authors’ names have been used as a full part of the name so
that, for example, van den Bergh, van Bibber, von Feilitzsch and others
are found under the letter V.
606
References 607
Anselm, A. A. 1982, Pis’ma Zh. Eksp. Teor. Fiz., 36, 46 (JETP Lett.,
36, 55).
Anselm, A. A. 1985, Yad. Fiz., 42, 1480 (Sov. J. Nucl. Phys., 42, 936).
Anselm, A. A. 1988, Phys. Rev. D, 37, 2001.
Anselm, A. A., and Uraltsev, N. G. 1982a, Phys. Lett. B, 114, 39.
Anselm, A. A., and Uraltsev, N. G. 1982b, Phys. Lett. B, 116, 161.
Arafune, J., and Fukugita, M. 1987, Phys. Rev. Lett., 59, 367.
Arafune, J., et al. 1987a, Phys. Rev. Lett., 59, 1864.
Arafune, J., et al. 1987b, Phys. Lett. B, 194, 477.
ARGUS Collaboration 1988, Phys. Lett. B, 202, 149.
ARGUS Collaboration 1992, Phys. Lett. B, 292, 221.
Arnett, W. D., and Rosner, J. L. 1987, Phys. Rev. Lett., 58, 1906.
Arnett, W. D., et al. 1989, Ann. Rev. Astron. Astrophys., 27, 629.
Ashkin, A. and Dziedzic, J. M. 1973, Phys. Rev. Lett., 30, 139.
Assamagan, K., et al. 1994, Phys. Lett. B, 335, 231.
Athanassopoulos, C., et al. 1995, Phys. Rev. Lett., 75, 2650.
Athar, H., Peltoniemi, J. T., and Smirnov, A. Yu. 1995, Phys. Rev. D,
51, 6647.
Atzmon, E., and Nussinov, S. 1994, Phys. Lett. B, 328, 103.
Aubourg, E., et al. 1993, Nature, 365, 623.
Aubourg, E., et al. 1995, Astron. Astrophys., 301, 1.
Auriemma, G., Srivastava, Y., and Widom, A. 1987, Phys. Lett. B, 195,
254.
Avignone, F. T., et al. 1987, Phys. Rev. D, 35, 2752.
Babu, K. S., Gould, T. M., and Rothstein, I. Z. 1994, Phys. Lett. B,
321, 140.
Babu, K. S., and Mohapatra, R. N. 1990, Phys. Rev. D, 42, 3866.
Babu, K. S., Mohapatra, R. N., and Rothstein, I. Z. 1992, Phys. Rev.
D, 45, R3312.
Babu, K. S., and Volkas, R. R. 1992, Phys. Rev. D, 46, R2764.
Bahcall, J. N. 1989, Neutrino Astrophysics (Cambridge University
Press).
Bahcall, J. N. 1990, Phys. Rev. D, 41, 2964.
Bahcall, J. N. 1991, Phys. Rev. D, 44, 1644.
Bahcall, J. N. 1994a, Phys. Rev. D, 49, 3923.
Bahcall, J. N. 1994b, Phys. Lett. B, 338, 276.
Bahcall, J. N. 1995, Neutrino 94, Nucl. Phys. B (Proc. Suppl.), 38, 98.
Bahcall, J. N., and Bethe, H. A. 1993, Phys. Rev. D, 47, 1298.
Bahcall, J. N., and Frautschi, S. C. 1969, Phys. Lett. B, 29, 623.
Bahcall, J. N., and Glashow, S. L. 1987, Nature, 326, 476.
Bahcall, J. N., and Holstein, B. R. 1986, Phys. Rev. C, 33, 2121.
References 609
Bahcall, J. N., Krastev, P. I., and Leung, C. N. 1995, Phys. Rev. D, 52,
1770.
Bahcall, J. N., and Pinsonneault, M. H. 1992, Rev. Mod. Phys., 64, 885.
Bahcall, J. N., and Pinsonneault, M. H. 1995, to be published in Rev.
Mod. Phys.
Bahcall, J. N., and Press, W. H. 1991, Ap. J., 370, 730.
Bahcall, J. N., and Ulrich, R. K. 1988, Rev. Mod. Phys., 60, 297.
Bahcall, J. N., and Wolf, R. A. 1965a, Phys. Rev. Lett., 14, 343.
Bahcall, J. N., and Wolf, R. A. 1965b, Phys. Rev., 5B, 1452.
Bahcall, J. N., et al. 1995, Nature, 375, 29.
Bailes, M. 1989, Ap. J., 342, 917.
Bailes, M., et al. 1990, Mon. Not. R. astr. Soc., 247, 322.
Bakalov, D., et al. 1994, Nucl. Phys. B (Proc. Suppl.), 35, 180.
Balantekin, A. B., and Loreti, F. 1992, Phys. Rev. D, 45, 1059.
Baluni, V. 1979, Phys. Rev. D, 19, 2227.
Balysh, A., et al. 1995, Phys. Lett. B, 356, 450.
Barbiellini, G., and Cocconi, G. 1987, Nature, 329, 21.
Barbieri, R., and Dolgov, A. 1991, Nucl. Phys. B, 349, 743.
Barbieri, R., and Fiorentini, G. 1988, Nucl. Phys. B, 304, 909.
Barbieri, R., and Mohapatra, R. N. 1988, Phys. Rev. Lett., 61, 27.
Barbieri, R., and Mohapatra, R. N. 1989, Phys. Rev. D, 39, 1229.
Barbieri, R., et al. 1991, Phys. Lett. B, 259, 119.
Bardeen, J., Bond, J., and Efstathiou, G. 1987, Ap. J., 321, 28.
Bardeen, W. A., Peccei, R. D., and Yanagida, T. 1987, Nucl. Phys. B,
279, 401.
Bardeen, W. A., and Tye, S. H. H. 1978, Phys. Lett. B, 74, 580.
Barger, V., Phillips, R. J. N., and Sarkar, S. 1995, Phys. Lett. B, 352,
365; (E) ibid., 356, 617.
Barger, V., Phillips, R. J. N., and Whisnant, K. 1991, Phys. Rev. D,
44, 1629.
Barger, V., Phillips, R. J. N., and Whisnant, K. 1992, Phys. Rev. Lett.,
69, 3135.
Baring, M. G. 1991, Astron. Astrophys., 249, 581.
Barnes, A. V., Weiler, T. J., and Pakvasa, S. 1987, Ap. J., 323, L31.
Barr, S. M., and Seckel, D. 1992, Phys. Rev. D, 46, 539.
Barroso, A., and Branco, G. C. 1982, Phys. Lett. B, 116, 247.
Barrow, J. D. 1978, Mon. Not. R. astr. Soc., 184, 677.
Barrow, J. D., and Burman, R. R. 1984, Nature, 307, 14.
Barstow, M. A. 1993, ed., White Dwarfs: Advances in Observation and
Theory (Kluwer, Dordrecht).
Battye, R. A., and Shellard, E. P. S. 1994a, Phys. Rev. Lett., 73, 2954.
610 References
Davis Jr., R., Mann, A. K., and Wolfenstein, L. 1989, Ann. Rev. Nucl.
Part. Sci., 39, 467.
Dearborn, D. S. P., Schramm, D. N., and Steigman, G. 1986, Phys.
Rev. Lett., 56, 26.
Dearborn, D. S. P., et al. 1990, Ap. J., 354, 568.
Debye, P., and Hückel, E. 1923, Phys. Z., 24, 185.
Degl’Innocenti, S., Fiorentini, G., and Lissia, M. 1995, Nucl. Phys. B
(Proc. Suppl.), 43, 66.
Degl’Innocenti, S., et al. 1995, Report MPI-PTh/95-78 and astro-ph/
9509090, submitted to Astron. Astrophys.
Degrassi, G., Sirlin, A., and Marciano, W. J. 1989, Phys. Rev. D, 39,
287.
del Campo, S., and Ford, L. H. 1988, Phys. Rev. D, 38, 3657.
Demarque, P., et al. 1994, Ap. J., 437, 870.
Dewey, R. J., and Cordes, J. M. 1987, Ap. J., 321, 780.
Dicus, D. A. 1972, Phys. Rev. D, 6, 961.
Dicus, D. A., Kolb, E. W., and Teplitz, V. L. 1977, Phys. Rev. Lett.,
39, 169.
Dicus, D. A., Kolb, E. W., and Tubbs, D. L. 1983, Nucl. Phys. B, 223,
532.
Dicus, D. A., and Repko, W. W. 1993, Phys. Rev. D, 48, 5106.
Dicus, D. A., et al. 1976, Ap. J., 210, 481.
Dicus, D. A., et al. 1978, Phys. Rev. D, 18, 1829.
Dicus, D. A., et al. 1980, Phys. Rev. D, 22, 839.
Dicus, D. A., et al. 1989, Phys. Lett. B, 218, 84.
DiLella, L. 1993, Neutrino 92, Nucl. Phys. B (Proc. Suppl.), 31, 319.
Dimopoulos, S., Starkman, G. D., and Lynn, B. W. 1986a, Phys. Lett.
B, 167, 145.
Dimopoulos, S., Starkman, G. D., and Lynn, B. W. 1986b, Mod. Phys.
Lett. A, 8, 491.
Dimopoulos, S., et al. 1986, Phys. Lett. B, 179, 223.
Dine, M., and Fischler, W. 1983, Phys. Lett. B, 120, 137.
Dine, M., Fischler, W., and Srednicki, M. 1981, Phys. Lett. B, 104, 199.
Dine, M., and Seiberg, N. 1986, Nucl. Phys. B, 273, 109.
Dirac, P. A. M. 1937, Nature, 139, 323.
Dirac, P. A. M. 1938, Proc. R. Soc., A165, 199.
Dobroliubov, M. I., and Ignatiev, A. Yu. 1990, Phys. Rev. Lett., 65,
679.
Dodd, A. C., Papageorgiu, E., and Ranfone, S. 1991, Phys. Lett. B,
266, 434.
Dodelson, S., and Feinberg, G. 1991, Phys. Rev. D, 43, 913.
References 617
Dodelson, S., Frieman, J. A., and Turner, M. S. 1992, Phys. Rev. Lett.,
68, 2572.
Dodelson, S., Gyuk, G., and Turner, M. S. 1994, Phys. Rev. Lett., 72,
3754.
Dolgov, A. D. 1981, Yad. Fiz., 33, 1309 (Sov. J. Nucl. Phys., 33, 700).
Dolgov, A. D., Kainulainen, K., and Rothstein, I. Z. 1995, Phys. Rev.
D, 51, 4129.
Dolgov, A. D., and Rothstein, I. Z. 1993, Phys. Rev. Lett., 71, 476.
Dolgov, A. D., and Raffelt, G. G. 1995, Phys. Rev. D, 52, 2581.
D’Olivo, J. C., Nieves, J. F., and Pal, P. B. 1989, Phys. Rev. D, 40,
3679.
D’Olivo, J. C., Nieves, J. F., and Pal, P. B. 1990, Phys. Rev. Lett., 64,
1088.
Domokos, G., and Kovesi-Domokos, S. 1995, Phys. Lett. B, 346, 317.
Donelly, T. W., et al. 1978, Phys. Rev. D, 18, 1607.
Dorofeev, O. F., Rodionov, V. N., and Ternov, I. M. 1985, Pis’ma As-
tron. Zh., 11, 302 (Sov. Astron. Lett., 11, 123).
Dydak, F., et al. 1984, Phys. Lett. B, 134, 281.
Dziembowski, W. A., et al. 1994, Ap. J., 432, 417.
Eggleton, P. P., and Cannon, R. C. 1991, Ap. J., 383, 757.
Eggleton, P. P., and Faulkner, J. 1981, in: Iben and Renzini (1981).
Ellis, J., and Karliner, M. 1995, Phys. Lett. B, 341, 397.
Ellis, J., and Olive, K. A. 1983, Nucl. Phys. B, 233, 252.
Ellis, J., and Olive, K. A. 1987, Phys. Lett. B, 193, 525.
Ellis, J., and Salati, P. 1990, Nucl. Phys. B, 342, 317.
Ellis, J., et al. 1986, Phys. Lett. B, 167, 457.
Ellis, J., et al. 1988, Phys. Lett. B, 215, 404.
Ellis, J., et al. 1989, Phys. Lett. B, 228, 264.
Engel, J., Krastev, P. I., and Lande, K. 1995, Report hep-ph/9501219,
submitted to Phys. Rev. D.
Engel, J., Seckel, D., and Hayes, A. C. 1990, Phys. Rev. Lett., 65, 960.
Enqvist, K., Rez, A. I., and Semikoz, V. B. 1995, Nucl. Phys. B, 436,
49.
Enqvist, K., and Uibo, H. 1993, Phys. Lett. B, 301, 376.
Ezer, D., and Cameron, A. G. W. 1966, Can. J. Phys., 44, 593.
Falk, S. W., and Schramm, D. N. 1978, Phys. Lett. B, 79, 511.
Faulkner, J., and Swenson, F. J. 1988, Ap. J., 329, L47.
Fayons, S. A., Kopeykin, V. I., and Mikaelyan, L. A. 1991, quoted after
L. Moscoso, Neutrino 90, Nucl. Phys. B (Proc. Suppl.), 19, 147
(1991).
Feinberg, E. L., and Pomeranchuk, I. 1956, Nuovo Cim. Suppl., 3, 652.
618 References
Goyal, A., Dutta, S., and Choudhury, S. R. 1995, Phys. Lett. B, 346,
312.
Greene, G. L., et al. 1991, Phys. Rev. D, 44, R2216.
Gregores, E. M., et al. 1995, Phys. Rev. D, 51, 4587.
Gribov, N. V., and Pontecorvo, B. M. 1969, Phys. Lett. B, 28, 493.
Grifols, J. A., and Massó, E. 1986, Phys. Lett. B, 173, 237.
Grifols, J. A., and Massó, E. 1989, Phys. Rev. D, 40, 3819.
Grifols, J. A., and Massó, E. 1990a, Phys. Lett. B, 242, 77.
Grifols, J. A., and Massó, E. 1990b, Nucl. Phys. B, 331, 244.
Grifols, J. A., Massó, E., and Peris, S. 1988a, Phys. Lett. B, 207, 493.
Grifols, J. A., Massó, E., and Peris, S. 1988b, Phys. Lett. B, 215, 593.
Grifols, J. A., Massó, E., and Peris, S. 1989a, Phys. Lett. B, 220, 591.
Grifols, J. A., Massó, E., and Peris, S. 1989b, Mod. Phys. Lett. A, 4,
311.
Grifols, J. A., Massó, E., and Peris, S. 1994, Astropart. Phys., 2, 161.
Grifols, J. A., Massó, E., and Rizzo, T. G. 1990, Phys. Rev. D, 42, 3293.
Grifols, J. A., and Tortosa, S. 1994, Phys. Lett. B, 328, 98.
Grimus, W., and Neufeld, H. 1993, Phys. Lett. B, 315, 129.
Guenther, D. B., et al. 1995, Ap. J., 445, 148.
Guzzo, M. M., Bellandi, J., and Aquino, V. M. 1994, Phys. Rev. D, 49,
1404.
Guzzo, M. M., Masiero, A., and Petcov, S. T. 1991, Phys. Lett. B, 260,
154.
Guzzo, M. M., and Petcov, S. T. 1991, Phys. Lett. B, 271, 172.
Guzzo, M. M., and Pulido, J. 1993, Phys. Lett. B, 317, 125.
Gvozdev, A. A., Mikheev, N. V., and Vassilevskaya, L. A. 1992a, Phys.
Lett. B, 289, 103.
Gvozdev, A. A., Mikheev, N. V., and Vassilevskaya, L. A. 1992b, Phys.
Lett. B, 292, 176.
Gvozdev, A. A., Mikheev, N. V., and Vassilevskaya, L. A. 1993, Phys.
Lett. B, 313, 161.
Gvozdev, A. A., Mikheev, N. V., and Vassilevskaya, L. A. 1994a, Phys.
Lett. B, 321, 108.
Gvozdev, A. A., Mikheev, N. V., and Vassilevskaya, L. A. 1994b, Phys.
Lett. B, 323, 179.
Haensel, P., and Jerzak, A. J. 1987, Astron. Astrophys., 179, 127.
Haft, M. 1993, Master’s Thesis, University of Munich, unpublished.
Haft, M., Raffelt, G., and Weiss, A. 1994, Ap. J., 425, 222; (E) 1995,
ibid., 438, 1017.
Hagmann, C., et al. 1990, Phys. Rev. D, 42, 1297.
Hagmann, C., and Sikivie, P. 1991, Nucl. Phys. B, 363, 247.
622 References
Totsuka, Y. 1993, Neutrino 92, Nucl. Phys. B (Proc. Suppl.), 31, 428.
Toussaint, D., and Wilczek, F. 1981, Nature, 289, 777.
Tsai, W., and Erber, T. 1975, Phys. Rev. D, 12, 1132.
Tsai, W., and Erber, T. 1976, Acta Phys. Austr., 45, 245.
Tsuruta, S. 1986, Comm. Astrophys., 11, 151.
Tsuruta, S. 1992, in: Pines, Tamagaki, and Tsuruta 1992.
Tsuruta, S., and Nomoto, K. 1987, in: A. Hewitt et al. (eds.), Obser-
vational Cosmology (IAU Sympsium No. 124).
Tsytovich, V. N. 1961, Zh. Eksp. Teor. Fiz., 40, 1775 (Sov. Phys. JETP,
13, 1249).
Tsytovich, V. N., et al. 1995, Phys. Lett. A, 205, 199.
Turck-Chièze, S., and Lopes, I. 1993, Ap. J., 408, 347.
Turck-Chièze, S., et al. 1993, Phys. Rep., 230, 57.
Turner, M. S. 1986, Phys. Rev. D, 33, 889.
Turner, M. S. 1987, Phys. Rev. Lett., 59, 2489.
Turner, M. S. 1988, Phys. Rev. Lett., 60, 1797.
Turner, M. S. 1992, Phys. Rev. D, 45, 1066.
Turner, M. S., Kang, H.-S., and Steigman, G. 1989, Phys. Rev. D, 40,
299.
Umeda, H., Nomoto, K., and Tsuruta, S. 1994, Ap. J., 431, 309.
Umeda, H., Tsuruta, S., and Nomoto, K. 1994, Ap. J., 433, 256.
Ushida, N., et al. 1986, Phys. Rev. Lett., 57, 2897.
Valle, J. W. F. 1987, Phys. Lett. B, 199, 432.
van Bibber, K., et al. 1987, Phys. Rev. Lett., 59, 759.
van Bibber, K., et al. 1989, Phys. Rev. D, 39, 2089
van Bibber, K., et al. 1992, Search for Pseudoscalar Cold Dark Matter,
Experimental Proposal, unpublished.
van Bibber, K., et al. 1994, Status of the Large-Scale Dark-Matter Axion
Search, Report UCRL-JC-118357 (Lawrence Livermore National
Laboratory).
van Buren, D., and Greenhouse, M. A. 1994, Ap. J., 431, 640.
VandenBerg, D. A., Bolte, M., and Stetson, P. B. 1990, Astron. J., 100,
445.
van den Bergh, S., and Tammann, G. A. 1991, Ann. Rev. Astron. As-
trophys., 29, 363.
van der Velde, J. C. 1989, Phys. Rev. D, 39, 1492.
van Horn, H. M. 1971, in: W. J. Luyten (ed.), White Dwarfs, IAU-
Symposium No. 42 (Reidel, Dordrecht).
Vassiliadis, G., et al. 1995, Proc. Int. Symp. Strangeness and Quark
Matter , Sept. 1–5, 1994, Crete, Greece (World Scientific, Singa-
pore).
640 References
642
Acronyms 643
L longitudinal
l.h. left-handed
l.h.s. left-hand side (of an equation)
ly light year
LMC Large Magellanic Cloud
LSP lightest supersymmetric particle
mfp mean free path
MS main sequence
MSW Mikheyev-Smirnov-Wolfenstein
NC neutral current
ND nondegenerate
T transverse
TL Turck-Chièze and Lopes
TO main-sequence turnoff
OPE one pion exchange
PQ Peccei-Quinn
QED quantum electrodynamics
QCD quantum chromodynamics
RG red giant
RGB red-giant branch
r.h. right-handed
r.h.s. right-hand side (of an equation)
SGB sub-giant branch
SN supernova
SNe supernovae
SNU solar neutrino unit
SNu supernova unit
SNBO Supernova Burst Observatory
UT universal time
VVO Voloshyn-Vysotskiı̆-Okun
WD white dwarf
Symbols
Symbols which have only a local meaning in, say, one paragraph are
not listed here. Four-momenta are usually denoted by uppercase italics
such as K, three-momenta are boldface lowercase letters such as k, the
modulus of a three-momentum is in lowercase italics such as k = |k|.
Lorentz indices are usually given in Greek letters (µ, ν, α, β, etc.),
three-indices and flavor indices in Latin letters (i, j, k, etc.).
Functions, Operators
Besides the usual functions and operators, the following convention may
be noteworthy.
Latin Symbols
a axion, axion field, annihilation operator, acceleration, radiation
constant (a = π 2 /15 in natural units)
ai ion-sphere radius
A electromagnetic vector potential, atomic mass number
B magnetic field, operator for “background” medium
Bω specific energy density of a radiation field
c speed of light (c = 1 in natural units)
cj speed of propagation of particle j
cp heat capacity at constant pressure
CV,A vector and axial-vector weak-coupling constants
Cj effective Peccei-Quinn charge of fermion j
d electric dipole moment
d, D distance
644
Symbols 645
Greek Symbols
α fine-structure constant (e2 /4π = 1/137), asymmetry parameter
in angular distribution of decay photons
α# fine-structure constant for general bosons, α# = g 2 /4π with the
Yukawa coupling g
αa,χ fine-structure constant for axions, majorons
αs strong fine-structure constant
απ pionic fine-structure constant, απ = (f 2mN /mπ )2 /4π ≈ 15 with
f ≈ 1.0
β velocity of a particle, parameter of axion models, parameter in
nucleon-nucleon-axion bremsstrahlung rate
γ photon, Dirac matrix, adiabatic index, dimensionless fluctuation
rate γ = Γ/T
Γ rate for decay or fluctuations, plasma coupling parameter
Γσ spin-fluctuation rate in a nuclear medium
δ δ function, Kronecker δ, differential quantity
* energy loss or generation rate per unit mass, electric dipole mo-
ment, polarization vector
*ij neutrino transition electric moment
η degeneracy parameter, η = (µ − m)/T
θ mixing angle, scattering angle
Θ Θ parameter of QCD
ΘW weak mixing angle (sin2 ΘW = 0.2325)
κ opacity, dimensionless screening or momentum scale
κ∗ reduced opacity
λ wave length, mean free path
µ magnetic moment, chemical potential, muon, Lorentz index,
mean molecular weight
µ̂ nonrelativistic chemical potential, µ̂ = µ − m
µij neutrino transition magnetic moment
648 Symbols
A direct search
decay photons of cosmic axions
α-Ori 189 491f
A665, A1413, A2218, A2256 (galaxy galactic (cavity search) 177f, 191f,
clusters) 490–92 544
adiabatic index 396 laboratory 182, 184f, 191f
adiabatic oscillations solar 100, 181, 191f, 462
→ neutrino oscillations SN 1987A 483
adiabatic temperature gradient 10–12 mass 527, 535
adiabaticity parameter 299 mixing with pion 527f, 534
models
AGB (asymptotic giant branch) 27, 35,
DFSZ 535–40
61
KSVZ 528–30, 542–44
aligned rotator 185
other 534–36
AMANDA xvi, 558
Nambu-Goldstone boson 190f, 528–30
anapole moment 268–70
quantum gravity 532f
Andromeda 446
solution to strong CP problem 527f
angular anomaly of SN 1987A neutrino
sphere 507
signal 422, 428–30
axio-recombination 100
antimatter supernova 497
arions 187–90
B
automatic symmetry 533
asymptotic giant branch 27, 35, 61 β decay 255
atmospheric neutrinos 290–3 ββ decay 257f, 560f
atomic mass unit 582, 591 Baade-Wesselink method 73
axio-electric effect 100 Baksan scintillator telescope (BST)
axion 414f, 417, 420
→ pseudoscalar boson band-structure effects 58, 106f
bounds 191f, 539–41 baryonic force 114
cavity experiment 177f, 191f, 544 beam-stop neutrinos 456f
cosmic density 491f, 528, 541–44 Betelgeuse 189
cosmic thermal production 541 big-bang nucleosynthesis: bounds
couplings majorons 259f, 561f
electron 536f millicharged particles 522f, 566f
fermions general 530–2 neutrino
gluon 159, 527, 530 Dirac dipole moment 277f
nucleon 536–38 mass 259f
photon 167f, 176, 191f, 536 time variation
decay 167f, 191f Fermi’s constant 546
decay constant 526f, 529, 535 Newton’s constant 547f
649
650 Subject Index
H I, J
Hayashi line 32, 35, 40 IMB detector
HB (horizontal branch) atmospheric neutrinos 291f
location in color-magnitude diagram SN 1987A neutrinos 415–23
27, 41, 61 inflation 542
morphology 34 ions
HB stars contribution to screening 220f, 596,
→ RR Lyrae stars 599
energy-loss argument 79–82, 98f, correlations 106, 222–25
107–9, 112, 176, 236 ion-sphere radius 223f, 597
inner structure 29f, 80f, 592, 599 isochrone 61f
lifetime 70, 76, 79–82 Jupiter magnetic field 556
overview 34f
particle bounds: tabulation 82 K
helioscope 179, 181f
helioseismology xvi, 17, 113, 344f, κ mechanism 40
549–51 Kamiokande (Cherenkov detector)
helium abundance atmospheric neutrinos 291f
presolar 16f, 351, 549f dark-matter search xvi
primordial 24, 67 neutrino detection
red-giant envelope 66f, 74–78 efficiency 417
helium-burning stars → HB stars electron scattering 365–67, 416f
helium flash → helium ignition proton absorption 416f
helium ignition proton decay xvii
→ red giant core mass SN 1987A neutrinos 418–23
brightness 63f, 68–72 solar neutrinos 343f, 368–73, 383
description 33f Kelvin-Helmholtz cooling
delay by particle emission 63, 83–87, Sun 8
236 supernova core 397, 400, 407–13,
off-center 34, 78, 83f 501–23
Hertzsprung gap 32, 40 Kibble mechanism 542
Hertzsprung-Russell diagram kick velocity (neutron stars) 443–45
→ color-magnitude diagram Klein-Gordon equation for mixed
Higgs field 3, 165f, 194, 252f, 528, 534, neutrinos 282–84, 294
545, 559 Kramers-Kronig relation 198
Homestake (solar neutrino detector) KSVZ axions 528–30, 542–44
chlorine 342f, 357–62, 371, 373–76
iodine 343, 394 L
homologous models 14–16, 549, 552
horizontal branch → HB Lamb shift 565f
hot bubble 38, 397, 406 Landau damping 199, 214, 216f
hot plus cold dark matter 258 Landau-Pomeranchuk-Migdal effect
Hulse-Taylor binary pulsar 145, 148f
→ PSR 1913+16 Landau-Zener approximation 300
Hyades 48 Langmuir wave
hydrodynamic event 6 → plasmon: longitudinal
hydrogen-burning chains 341, 346f Large Magellanic Cloud 39, 74, 414, 446
hydrostatic equilibrium 5–7 large-numbers hypothesis 545, 549
654 Subject Index