Stars As Laboratories For Fundamental Physics
Stars As Laboratories For Fundamental Physics
Stars As Laboratories For Fundamental Physics
Raelt
Stars as Laboratories
for Fundamental Physics
The Astrophysics of Neutrinos, Axions, and Other
Weakly Interacting Particles
Table of Contents
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)
vi
Table of Contents
Table of Contents
vii
viii
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).
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 Oscillations (284) 3. Distribution of Sources and Energies (287) 4. Experimental
Oscillation Searches (289) 5. Atmospheric Neutrinos (290)
3. Weak-Damping
Table of Contents
Table of Contents
xi
xii
Table of Contents
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)
15.3
15.4
15.5
15.6
15.7
Table of Contents
xiii
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 Process (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 Newtons and Einsteins theories of gravity.
This tradition carries onthe 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 superlunar sphere has expanded far beyond these traditional elds. 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 gravitational physics and the nature of space and time (Is Newtons constant
constant? Do all relativistic particles move with the same limiting velocity? Are there novel long-range interactions?), most of the discussion
focusses on the properties and nongravitational interactions of elementary particles.
There are three predominant methods for the use of stars as particlephysics 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 astronomical distances before reaching the detector one can study modications
of the measured signal which can be attributed to propagation and dispersion eects, including neutrino avor oscillations or axion-photon
oscillations in intervening magnetic elds. It is well known that the
discrepancy between the calculated and measured solar neutrino spectra 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 -rays yields a limit on neutrino radiative 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 -rays in conjunction with the SN 1987A neutrinos which allows one to conclude, for example, that even must obey
the cosmological limit of m <
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 eect
has been routinely included in stellar evolution calculations. If new lowmass elementary particles were to exist such as axions or other NambuGoldstone 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 results 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 literature 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 topical conferences (Madsen and Haensel 1992; Vassiliadis et al. 1995).
I have also dodged some important issues in the three-way relationship between cosmology, stars, and particle physics. If axions are not
the dark matter of the universe, it is likely lled with a background
sea of hypothetical weakly interacting massive particles (WIMPs) such
as the lightest supersymmetric particles. Moreover, there may be exotic particles left over from the hot early universe such as magnetic
monopoles which are predicted to exist in the framework of typical
grand unied 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-Callaneect), 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 example, 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 because their mean free path can be so large that they may be orbiting almost freely in the stars gravitational potential well, with only
occasional collisions with the background medium. Originally it was
thought that this eect 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 mechanism now seems to cause a discrepancy with the observed solar p-mode
frequencies (Christensen-Dalsgaard 1992). In addition, a signicant effect 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 terrestrial detectors such as Kamiokande or the future Superkamiokande,
NESTOR, DUMAND, and AMANDA Cherenkov detectors. This indirect approach to search for dark matter may well turn into a serious
competitor for the new generation of direct laboratory search experiments that are currently being mounted. This material is extensively
covered in a forthcoming review Supersymmetric Dark Matter by Jungman, Kamionkowski, and Griest (1995); there is no need for me to
duplicate the eort of these experts.
The topics covered in my book revolve around the impact of lowmass or massless particles on stars or the direct detection of this radiation. 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 laboratories for fundamental physics help to push the low-energy frontier of
particle physics and as such complement the eorts of nonaccelerator
particle experiments. Their main thrust is directed at the search for
nonstandard neutrino properties, but there are other fascinating topics 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 Frejus detector, 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 nd proton decay.
xviii
Preface
The complete or near masslessness of the particles studied here (neutrinos, photons, axions, etc.) opens up a rich phenomenology in its own
right. One intriguing issue is the production and propagation of these
objects in a hot and dense medium which modies their dispersion relations in subtle but signicant ways. One of the most important neutrino
production processes in stars is the photon decay which is
enabled by the modied photon dispersion relation in a medium, and
by an eective neutrino-photon coupling mediated by the ambient electrons. Another example is the process of resonant neutrino oscillations
(MSW eect) which is instrumental at explaining the measured solar
neutrino spectrum, and may imply vast modications of SN physics,
notably the occurrence of r-process nucleosynthesis. The MSW eect
depends on a avor-birefringent term of the neutrino dispersion relation in a medium. In a SN, neutrinos themselves make an important
contribution to the background medium so that their oscillations become a nonlinear phenomenon. In the deep interior of a SN core where
neutrinos are trapped, a kinetic treatment of the interplay between oscillations and collisions requires a fascinating nonabelian Boltzmann
collision equation.
One may think that low-energy particle physics in a dense and hot
medium requires the tools of eld theory at nite temperature and
density (FTD) which has taken a stormy development over the past
fteen years. In practice, its contribution to particle astrophysics has
been minor. The derivation of dispersion relations in a medium can be
understood in kinetic theory from forward scattering on the medium
constituentsdispersion is a lowest-order phenomenon. The imaginary
part of a particle self-energy in the medium is physically related to its
emission and absorption rate which in FTD eld theory is given by
Preface
xix
xx
Preface
scales that have been used for the purposes of particle astrophysics.
For each case the salient applications are summarized. Chapters 36
deal with the interaction of radiation (neutrinos, axions, other lowmass 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 justies this arrangement.
Chapter 6 develops the topic of particle dispersion in media. The
medium-induced photon dispersion relation allows for the plasma process 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 argument and the dispersion arguments of the following chapters.
Dispersion and propagation eects are particularly important for
massive neutrinos with avor mixing. To this end the phenomenology
of massive, mixed neutrinos is introduced in Chapter 7. Vacuum and
matter-induced avor oscillations as well as magnetically induced spin
oscillations are taken up in Chapter 8. If neutrinos are in thermal equilibrium as in a young SN core or the early universe, neutrino oscillations
require a dierent theoretical treatment (Chapter 9).
Chapters 1013 are devoted to astrophysical sources where neutrinos have been measured, i.e. the Sun and supernovae (for the latter 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 gure 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 1416 give particle-specic 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 boson for which they serve as a generic example. In Chapter 14 these
results are interpreted in terms of axion-specic 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
Parts of my presentation are devoted to theoretical and calculational ne points of particle dispersion and emission eects in media. I
nd some of these issues quite intriguing in their own right. Still, the
main goal has been to provide an up-to-date overview of what we know
about elementary particles and their interactions on the basis of established stellar properties and on the basis of measured or experimentally
constrained stellar particle uxes.
All those whose lives are spent searching for truth are well
aware that the glimpses they catch of it are necessarily eeting, glittering for an instant only to make way for new and
still more dazzling insights. The scholars 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, 18621935)
In spite of this bittersweet insight, and in spite of some inevitable
errors of omission and commission, I hope that my book will be of some
use to researchers, scholars, and students interested in the connection
between fundamental physics and stars.
Munich, May 1995.
Acknowledgments
While writing this book I have benetted from the help and encouragement of many friends and colleagues. In particular, I need to mention
Hans-Thomas Janka, Lothar Luh, G
unter 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 eort 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 BerkeleyI gratefully acknowledge the ne hospitality of Bernard Sadoulet and his sta. Staying for that period as a guest in Edward Janellis house made a huge
dierence. 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 transform my manuscript into a book. David Schramm as the series editor
originally solicited this opus (just expand your Physics Report a little
bit). It hasnt quite worked that way, but now that Im nished I am
grateful that David persuaded me to take up this project.
xxii
Chapter 1
The Energy-Loss Argument
Weakly interacting, low-mass particles such as neutrinos or axions contribute to the energy loss or energy transfer in stars. The impact of
an anomalous energy-loss mechanism is discussed qualitatively and in
terms of homology relations between standard and perturbed stellar
models. The example of massive pseudoscalar particles is used to illustrate the impact of a new energy-loss and a new radiative-transfer
mechanism on the Sun.
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 Bethes (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 reactions of the sort e + (A, Z)
(A, Z1) + e or (A, Z1) (A, Z) + e + e , the so-called URCA
reactions which Gamow and Schoenberg had in mind, or by fusion processes like pp de+ e . 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
Chapter 1
Chapter 1
1.2
1.2.1
Chapter 1
(1.1)
(1.2)
This is the virial theorem which is the most important tool to understand the behavior of self-gravitating systems.
As a simple example for the beauty and power of the virial theorem one may estimate the solar central temperature from its mass and
radius. The material is dominated by protons which have a gravitational potential energy of order GN M mp /R = 2.14 keV where
M = 1.991033 g is the solar mass, R = 6.961010 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 = 0.8107 K. This is to be compared with 1.56107 K
found for the central temperature of a typical solar model. This example illustrates that the basic properties of stars can be understood from
simple physical principles.
1.2.2
a) Normal Stars
There are two main sources of pressure relevant in stars, thermal pressure and degeneracy pressure. The third possibility, radiation pressure,
never dominates except perhaps in the most massive stars. The pressure provided by a species of particles is proportional to their density,
to their momentum which is reected 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 (/) T with the mass density and the mean molecular weight of the medium constituents. For nonrelativistic degenerate
electrons the density is ne = p3F /3 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
and thus to 5/3 .
e
The two main pressure sources determine two generic forms of behavior 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 dierent logic.
A normal star is understood most easily if one imagines how it initially 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 conguration, the virial theorem Eq. (1.2) in-
Chapter 1
The most salient feature of a normal stellar conguration is the interplay of its negative specic heat and nuclear energy generation. Conversely, if the pressure were dominated by electron degeneracy it would
be nearly independent of the temperature. Then this self-regulation
would not function because heating would not lead to expansion. Thus,
stable nuclear burning and the dominance of thermal pressure go inseparably hand in hand. Another salient feature of such a conguration is
the inevitability of its nal demise because it lives on a nite supply of
nuclear fuel.
b) Degenerate Stars
Everything is dierent for a conguration 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 actually 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 conditions 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 conguration 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 5/3 is some average pressure. Because
M/R3 one nds p/R M5/3 /R6 while the r.h.s. of Eq. (1.1) is
proportional to M/R2 and thus to M2 /R5 . Therefore, a degenerate
conguration is characterized by R M1/3 .
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 xed at c (or 1 in natural units), causing the pressure
to vary only as p4F or 4/3 . In this case adding mass no longer leads to a
sucient pressure increase to balance for the extra weight. Beyond this
Chandrasekhar limit, which is about 1.4 M for a chemical composition with Ye = 21 (number of electrons per baryon), no stable degenerate
conguration exists.
In summary, the salient features of a degenerate conguration are
the inverse mass-radius relationship R M1/3 , the Chandrasekhar
limit, the absence of nuclear burning, and the positive specic heat
which allows the conguration to cool when it loses energy.
10
Chapter 1
c) Giant Stars
A real star can be a hybrid conguration 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 conguration. The envelope follows the self-regulating
logic of a normal star except that it is no longer dominated by its selfgravity 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. Lowmass red giants will play a major role in Chapter 2a further discussion
of their fascinating story is deferred until then.
1.2.3
Energy Conservation
(1.3)
(1.4)
(1.5)
For recent attempts see Faulkner and Swenson (1988), Eggleton and Cannon
(1991), and Renzini et al. (1992).
11
Energy Transfer
4r2 d(aT 4 )
,
3
dr
(1.6)
(1.7)
12
Chapter 1
13
>
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
aects the composition prole 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 conguration. 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 advancing hydrogen-burning shell which encounters a discontinuity in the
hydrogen prole. 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
identied in several clusters (Figs. 2.18 and 2.19); its location is in good
agreement with theoretical expectations (Fusi Pecci et al. 1990).
1.2.5
Gravitational Settling
The composition prole of a star can also change by diusion, and notably by gravitational settling of the heavier elements. This eect was
ignored in most evolution calculations because the time scales are very
large. Still, the settling of helium will displace hydrogen from the center of a hydrogen-burning star and thus accelerate the depletion and
main-sequence turno. The gravitational settling of metals will lead
to an opacity increase in the central regions. Helium settling reduces
the inferred globular cluster ages by 1 Gyr or more which is about a
10% eect (Prott and Michaud 1991; Chaboyer et al. 1992). Because the inferred globular cluster ages are larger than the expansion
14
Chapter 1
1.3
1.3.1
15
and
s T p .
(1.8)
For the opacity, Frieman et al. took the Kramers law with s = 1
and p = 3.5 which is found to be a reasonable interpolation formula throughout most lower main-sequence interiors. Hence, the local
energy ux scales as
L (r ) = y 1/2 L(r) .
(1.9)
The hydrogen-burning rate nuc also has the required form with n = 1,
and for the pp chain = 46; it dominates in the Sun and in stars
with lower mass. It is assumed that the new energy-loss rate x follows
the same proportionality; the standard neutrino losses are ignored
because they are small on the lower main sequence. If the star is not
in a phase of major structural readjustment one may also ignore grav
in Eq. (1.4) so that in Eq. (1.3)
= (1 x ) nuc ,
(1.10)
(1.11)
leading to
y = (1 x )2/(2+5) .
(1.12)
16
Chapter 1
even if the luminosity Lx in exotics is as large as the photon luminosity (x = 12 ) the overall changes in the stellar structure remain moderate. The predominant eect is an increased consumption of nuclear
fuel at an almost unchanged stellar structure, leading to a decreased
duration of the hydrogen-burning phase of
/ x .
(1.14)
(1.15)
L
5x
=
,
L
2 + 11
T
2x
=
.
T
2 + 11
(1.16)
These stars also contract, and the internal temperature increases, but
the surface luminosity decreases.
1.3.2
For the Sun, the radius and luminosity are very well measured and
so one may think that small deviations R and L from a standard
model were detectable. This is not so, however, because a solar model
is defined to produce the observed radius and luminosity at an age
of 4.5 Gyr. The unknown presolar helium abundance Yinitial is chosen
to reproduce the present-day luminosity, and the one free parameter
of the mixing-length theory relevant for superadiabatic convection is
calibrated by the solar radius.
In a numerical study Raelt and Dearborn (1987) implemented axion losses by the Primako process in a 1 M stellar model, metallicity Z = 0.02, which was evolved to 4.5 Gyr with dierent amounts
of initial helium and dierent axion coupling strengths. Details of the
emission rate as a function of temperature and density are studied in
17
Table 1.1. Initial helium abundance for solar models with axion losses.
g10
Yinitial
Xc
0
10
15
20
25
0.274
0.266
0.256
0.241
0.224
0.00
0.16
0.32
0.51
0.65
0.362
0.307
0.292
0.245
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 ga .
Without exotic losses a presolar helium abundance of Yinitial = 0.274
was needed to reproduce the present-day Sun. For several values of
g10 ga /1010 GeV1 Raelt and Dearborn found the initial helium
values given in Tab. 1.1 necessary to produce the present-day luminosity. The values for x in Tab. 1.1 are dened as in Eq. (1.15) with Lx
the axion luminosity of the (perturbed) present-day solar model which
has L = L . Also, the central hydrogen abundance Xc of the presentday model is given. For g10 = 30, corresponding to x 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 x less than
about 0.5 is hard to exclude on the basis of this calculation. Therefore,
1
<
the approximate solar constraint remains x <
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 excellent agreement between standard solar models and the observed p-mode
frequencies there is little leeway for a modied solar structure and composition. This method has been used to constrain a hypothetical time
variation of Newtons constant (Sect. 15.2.3).
1.3.3
If novel particles are so weakly interacting that they escape freely from
the star once produced their role is that of a local energy sink. Neutrinos are of that nature, except in supernova cores where they are
trapped for several seconds. One could imagine new particles with
18
Chapter 1
such large interactions that they are even trapped, say, in the Sun.
Because their mean free path (mfp) is now less than the geometric dimension 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 opacity. 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 dierential equation (1.6). In
this case the dierence between an energy-loss and an energy-transfer
mechanism is blurred.
Equation (1.6) is justied when the mfp is less than the temperature scale height (d ln T /dr)1 . For radiative transfer ()1 is an
average mfp. Therefore, Eq. (1.6) informs us that the energy ux carried 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 temperature gradient. Radiative transfer is more ecient for a larger mfp, i.e.
for a more weakly interacting particle!
If a second photon existed with a coupling strength instead of
1
= 137
, it would contribute more to the energy transfer for < .
The observed properties of the Sun and other stars conrm 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 L is less than the photon luminosity L .
1.3.4
In order to implement the energy-transfer argument one needs a properly dened expression for the Rosseland opacity contribution of arbitrary bosons. Following the textbook derivation of the photon radiative
opacity such a denition was provided by Carlson and Salati (1989) as
well as Raelt and Starkman (1989).
For a suciently short mean free path the radiation eld of the
new bosons is taken to be locally isotropic. The local energy ux is then
found to be F = 31 B where the index indicates that a quantity refers to the boson energy . Moreover, B and F are understood
to be specic, i.e. an energy density and ux per unit energy interval.
The velocity is = [1 (m/)2 ]1/2 for a boson with mass m. (Recall
that natural units with h
= c = kB = 1 are always used.) In local
thermal equilibrium the energy density for massive bosons correspond-
19
g 2 ( 2 m2 )1/2
.
2 2
e/T 1
(1.17)
d T B ,
(1.18)
d T B .
x
4aT 3 m
(1.19)
The exotic opacity thus dened appears in the stellar structure equation in the way indicated by Eq. (1.7).
The production and absorption of bosons involves a Bose stimulation factor. This eect is taken account of by including a factor
(1 e/T ) under the integral in Eq. (1.19). The absorptive opacity
thus derived is usually referred to as the reduced opacity which is
the quantity relevant for energy transfer. In practice, it is not very
dierent from as is typically 3T for massless bosons.
In the large-mass limit (m T ) the reduction factor may be ignored
entirely and one nds to lowest order,
1
15
=g 4
x
4
m
T
)3
em/T
dy (y) y ey ,
(1.20)
20
1.3.5
Chapter 1
Solar Bound on Massive Pseudoscalars
160x Ye T 6
F (mx /T ),
mu m4e
(1.21)
Fig. 1.1. Eects of massive pseudoscalar particles on the Sun (interior temperature about 1 keV). Above the dashed line they contribute to the radiative energy transfer, below they escape freely and drain the Sun of energy.
The shaded area is excluded by this simple argument. (Adapted from Raelt
and Starkman 1989.)
21
scalars of
2(2)9/2 x Ye T 5/2 emx /T
x =
1/2
45
mx m4e
2.5 mx /T
,
= 4.4103 cm2 g1 x Ye m0.5
keV TkeV e
(1.22)
1.4
General Lesson
22
Chapter 1
Chapter 2
Anomalous
Stellar Energy Losses
Bounded by Observations
After a description of the main phases of stellar evolution, a review is
given of the observations that have been used to constrain novel stellar
energy losses. The main arguments involve the cooling speed of white
dwarfs and old neutron stars, the delay of helium ignition in low-mass
red giants, and the helium-burning lifetime of horizontal-branch stars.
The latter two arguments, which are based on observations of globularcluster stars, are cast into the form of an easy-to-use analytic criterion
that allows for a straightforward application in many dierent cases.
The cooling speed of nascent neutron stars is discussed in Chapter 11
in the context of supernova neutrinos.
2.1
2.1.1
24
Chapter 2
An excellent starting point for the nonexpert is Shu (1982). The classics are
Chandrasekhar (1939), Schwarzschild (1958), Clayton (1968), and Cox and Giuli
(1968). A recent general textbook is Kippenhahn and Weigert (1990). Recent monographs specializing on the Sun in general are Stix (1989), and on solar neutrinos
Bahcall (1989). The theory of stellar pulsation is covered in Cox (1980). A classic on
the physics of compact stars is Shapiro and Teukolsky (1983). A recent monograph
on neutron stars is Lipunov (1992) and on pulsars Meszaros (1992). Recent collections of papers or conference proceedings are available on the physics of red giants
(Iben and Renzini 1981), white dwarfs (Barstow 1993), pulsating stars (Schmidt
1989), neutron stars (Pines, Tamagaki, and Tsuruta 1992), and supernovae (Brown
1988; Petschek 1990). Many excellent reviews are found in the Annual Review of
Astronomy and Astrophysics.
25
26
Chapter 2
gas; then no further star formation is possible.4 Therefore, these systems are very clean laboratories for studies of stellar evolution as they
contain an early generation of stars (typical metallicities 103 104 )
which in a given cluster are all coeval and have almost equal chemical
compositions. To a rst approximation the stars in a globular cluster
dier only in one parametertheir initial mass.
The virial theorem explains (Sect. 1.2.2) that the negative specic
heat of a protostellar cloud leads to contraction and heating until it has
reached a thermal equilibrium conguration where the nuclear burning
rate exactly balances its overall luminosity. At this point the conguration is determined entirely by its mass, apart from a small inuence
of the metal content. Therefore, after the initial contraction hydrogenburning stars of dierent mass form the zero-age main sequence in
a Hertzsprung-Russell diagram where the eective surface temperature
is plotted on the horizontal axis and the stellar luminosity on the vertical axis (Fig. 2.2). As hydrogen is consumed small adjustments of the
conguration occur. However, essentially a star remains at its initial
main-sequence location for most of its life.
The escape velocity from our galaxy is about 500 km s1 . However, in a galactic
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.
4
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 classication has been
adopted for the evolutionary phases. MS (main sequence): core hydrogen
burning. BS (blue stragglers). TO (main-sequence turno): central hydrogen 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): helium and hydrogen shell burning. P-AGB (post-asymptotic giant branch):
nal evolution from the AGB to the white-dwarf stage. (Original of the
gure courtesy of A. Renzini.)
28
Chapter 2
The brightness of a star is photometrically measured on the basis of the luminosity in a wavelength band dened by a lter with a well-dened spectral response.
There exist many dierent color systems, corresponding to many dierent lters. In
Fig. 2.3 the visual brightness is dened by V = 2.5 log LV + const. with LV the luminosity measured with the visual lter, centered in the yellow-green waveband,
and a similar denition applies to the brightness B in the blue. The color B V
is a measure of the surface temperature with lower temperatures (redder color) to
the right. For technical denitions see Allen (1963). Note that the downward turn
of the horizontal branch in the blue is determined by the V lter; bolometrically
the HB is truly horizontal. An absolute bolometric brightness is given by
Mbol = 4.74 2.5 log(L/L ).
It is a dimensionless number, but often the unit mag (magnitude) is appended for
clarity. The visual brightness in Fig. 2.3 is given in apparent magnitudes which depend on the distance of the cluster. However, because of the logarithmic brightness
denition a dierent distance leads only to a vertical shift of the entire picture; the
color and relative brightnesses remain unchanged.
29
Fig. 2.4. Left panels: A typical solar model (Bahcall 1989) as an example 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
2.5107 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.
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 Dearborns evolution code.)
32
Chapter 2
33
Helium Ignition
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 particles (He nuclei) is not stable, He burning proceeds
directly to carbon, 3 12 C (triple- reaction), via an intermediate
8
Be state. Because it is essentially a three-body reaction its rate depends sensitively on and T . In a red-giant core helium ignites when
Mc 0.5 M with central conditions of 106 g cm3 and T 108 K
where the triple- energy generation rate per unit mass, 3 , varies approximately as 2 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 3 exceeds
the neutrino losses a nuclear runaway occurs. Because the pressure
is mainly due to degenerate electrons the energy production at rst
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 specic heat kicks in (Chapter 1). The explosion
energy is absorbed by the work necessary to expand the core from
about 106 g cm3 to about 104 g cm3 . The core temperature of the nal
conguration remains at about 108 K because of the steep temperature
dependence of 3 which allows only for a narrow range of stationary
burning conditions.
The nal conguration with a helium-burning core and a hydrogenburning shell (Fig. 2.5) is known as a horizontal-branch (HB) star, a
term which is justied by the location of these objects in the colormagnitude diagram Fig. 2.3. Note that the overall luminosity has decreased 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 ash. 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
For recent synthetic HBs see Lee, Demarque, and Zinn (1990, 1994).
35
remaining for an HB star after mass loss on the RGB will still depend
on its initial value, rendering this a rather natural possibility. However, it implies that the globular clusters of our galaxy did not form
at the same timean 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 xed 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.
2.1.5
After the exhaustion of helium at the center a degenerate carbonoxygen (CO) core forms with helium shell burning and continuing hydrogen shell burning (Fig. 2.5). Again, the star grows progressively
brighter and inates. Put another way, it becomes very similar to a
star which rst ascended the RGB: it ascends its Hayashi line for a
second time. The track in the Hertzsprung-Russell diagram asymptotically approaches that from the rst ascent (asymptotic giants).
The upper RGB can be observationally dicult to distinguish from the
asymptotic giant branch (AGB) even though they are reasonably well
separated in Fig. 2.3.
In low-mass stars carbon and oxygen never ignite. The shell sources
extinguish when most of the helium and hydrogen has been consumed
so that the star has lost its entire envelope either by hydrogen burning
or by further mass loss on the AGB. The remaining degenerate CO
star continues to radiate the heat stored in its interior. At rst these
stars are rather hot, but geometrically very small with a typical radius of 104 km. Their small surface area restricts their luminosity in
spite of the high temperature and so they are referred to as white
dwarfs. Because they are supported by electron degeneracy pressure,
their remaining evolution is cooling by neutrino emission from the interior and by photon emission from the surface until they disappear
from visibility.
36
Chapter 2
Fig. 2.7. Ring Nebula in Lyra (M57), a planetary nebula. (Image courtesy
of Palomar/Caltech.)
For stars which start out suciently 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).
2.1.6
Type I Supernovae
37
Intermediate-Mass Stars
The course of evolution for massive stars (M > 68 M ) is qualitatively dierent because they ignite carbon and oxygen in their core,
allowing them to evolve further. This is possible because even after
mass loss they are left with enough mass that their CO core grows
toward the Chandrasekhar limit. Near that point the density is high
enough to ignite carbon which causes heating and thus temporarily relieves the electron degeneracy. Next, the ashes of carbon burning (Ne,
Mg, O, Si) form a degenerate core which ultimately ignites Ne burning,
and so forth. Ultimately, the star has produced a degenerate iron core,
surrounded by half a dozen onion rings of dierent burning shells.
The game is over when the iron core reaches its Chandrasekhar limit
because no more nuclear energy can be released by fusion. The temperature is at 0.81010 K = 0.7 MeV, the density at 3109 g cm3 , and
there are about Ye = 0.42 electrons per baryon. Further contraction
leads to a negative feedback on the pressure as photons begin to dissociate iron, a process which consumes energy. Electrons are absorbed
and converted to neutrinos which escape, lowering the electron Fermi
momentum and thus the pressure. Therefore, the core becomes unstable and collapses, a process which is intercepted only when nuclear
density (31014 g cm3 ) is reached where the equation of state stiens.
38
Chapter 2
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 reected and thus turned into
an explosion. In practice, it is dicult 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
elds (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),
39
that even neutrinos are trapped. Therefore, energy and lepton number are lost approximately on a neutrino diusion time scale of several
seconds. The neutrinos from stellar collapse were observed for the rst
and only time when the star Sanduleak 69 202 in the Large Magellanic 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 horizontally across the Hertzsprung-Russell diagram until they reach their
Hayashi line, i.e. until they have become red supergiants. However, subsequently 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 ones lifetime. The visible galactic SNe previous to
1987A were Tychos and Keplers in close succession about 400 years
ago. Both of them may have been of type Ino 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.
2.1.9
Variable Stars
40
Chapter 2
41
42
2.2
2.2.1
Chapter 2
White-Dwarf Cooling
Theoretical and Observed White-Dwarf Properties
After this general survey of how stars evolve it is time to study individual aspects in more detail, and notably, how stellar evolution is aected
by the emission of weakly interacting particles. I begin with the conceptually most transparent case of stars for which the loss of energy
simply accelerates their cooling, i.e. white dwarfs and neutron stars.
The former represent the nal state of the evolution of stars with
initial masses of up to several M , perhaps up to 8 M (Sect. 2.1).
For reviews of the theory and observed properties see Hubbard (1978),
Liebert (1980), Shapiro and Teukolsky (1983), Weidemann (1990), and
DAntona and Mazzitelli (1990). Because white dwarfs (WDs) are supported by electron degeneracy pressure the hydrostatic and thermal
properties are largely decoupled. In Sect. 1.2.2 we had encountered
their inverted mass-radius relationship; in a polytropic approximation
of the WD structure one nds quantitatively (Shapiro and Teukolsky 1983)
R = 10,500 km (0.6M /M)1/3 (2/e )5/3 .
(2.1)
1
Here, M is the stellar mass and e = m1
= Ye1 the mean
u ne
molecular weight of the electrons with the mass density, mu the
atomic mass unit, ne the electron density, and Ye the number of electrons per baryon. WDs do not contain any hydrogen in their interiorit
would immediately igniteso that e = 2. Typically they consist of
carbon and oxygen, the end products of helium burning in the core of
the progenitor star. The central density of a polytropic model is
(2.2)
(2.3)
43
asymptotic giant evolution which can amount to an ejection of the entire envelope and thus to the formation of a planetary nebula (Fig. 2.7).
A theoretical evolutionary track for a 3 M star from the MS to the
WD stage was calculated, e.g. by Mazzitelli and DAntona (1986). The
central stars of planetary nebulae are identied with nascent WDs. The
rate of WD formation inferred from the luminosity function discussed
below agrees within a factor of about 2 with the observed formation
rate of planetary nebulae, which means that both quantities agree to
within their statistical and systematic uncertainties.
The hottest and brightest WDs have luminosities of L 0.5 L
while the faintest ones are observed at L 0.5104 L . Thus, because of their small surface area WDs are intrinsically faint (see the
Hertzsprung-Russell diagram Fig. 2.9). This implies that they can be
observed only in the solar neighborhood, for bright WDs out to about
100 pc (300 lyr). Because their vertical scale height in the galactic disk
is about 250 pc (Fleming, Liebert, and Green 1986) the observed WDs
homogeneously ll a spherical volume around the Sun. One may then
express the observed number of WDs in terms of a volume density; it
is of order 102 pc3 .
The observed luminosity function (the space density of WDs per
brightness interval) is shown in Fig. 2.10 and listed in Tab. 2.1 according
to Fleming, Liebert, and Green (1986) and Liebert, Dahn, and Monet
(1988). The operation of a novel cooling mechanism can be constrained
by three important features which characterize the luminosity function:
its slope, which signies the form of the cooling law, its amplitude,
which characterizes the cooling time and WD birthrate, and its sudden
break at log(L/L ) 4.7, which characterizes the beginning of WD
formation. Even the oldest WDs have not had time to cool to lower
luminosities. From this break one can infer an age for the galactic
disk of 810.5 Gyr (Winget et al. 1987; Liebert, Dahn, and Monet
1988; Iben and Laughlin 1989; Wood 1992) while Hernanz et al. (1994)
nd 9.512 Gyr on the basis of their cooling calculations which include
crystallization eects.
Because of the WD mass-radius relationship the surface temperature and luminosity are uniquely related for a given WD mass. Therefore, instead of the luminosity function one may consider the temperature distribution which, in principle, is independent of uncertain WD
distance determinations. Fleming, Liebert, and Green (1986) gave the
distribution of their sample of hot WDs in several temperature bins
(Tab. 2.2). Numerical cooling calculations of Blinnikov and DuninaBarkovskaya (1994) found good agreement with this distribution, as-
44
Chapter 2
MV
9.5
10.0
10.5
11.0
11.5
12.0
12.5
13.0
Mean
Mbol
Mean
log(L/L )
dN/dMbol
[pc3 mag1 ]
5.50
6.88
7.84
8.92
10.12
11.24
11.98
12.55
13.50
14.50
15.50
0.31
0.86
1.25
1.68
2.16
2.61
2.90
3.13
3.51
3.91
4.31
1.22106
1.01105
2.16105
9.56105
1.21104
1.51104
2.92104
6.07104
0.89103
1.34103
0.24103
log(dN/dMbol )
5.91
5.00
4.67
4.02
3.92
3.82
3.54
3.22
3.05
2.87
3.62
(+0.18,0.31)
(+0.14,0.21)
(+0.13,0.18)
(+0.12,0.16)
(+0.11,0.15)
(+0.11,0.16)
(+0.11,0.16)
(+0.20,0.39)
(+0.14,0.21)
(+0.14,0.20)
(+0.18,0.31)
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 dierence is taken as a conrmation of the standard WD cooling theory.
45
Fig. 2.10. Observed WD luminosity function as in Tab. 2.1. The dotted line represents Mestels cooling law with a constant WD birthrate of
B = 103 pc3 Gyr1 . The dashed line is from the numerical cooling curve
of a 0.6 M WD (Koester and Schonberner 1986), including neutrino losses
and assuming the same constant birthrate.
2.2.2
Cooling Theory
A WD has no nuclear energy sources and so it shines on its residual thermal energy: the evolution of a WD must be viewed as a cooling process
(Mestel 1952). Because electron conduction is an ecient mechanism
of energy transfer the interior can be viewed, to a rst approximation,
as an isothermal heat bath with a total amount of thermal energy U .
Because the nondegenerate surface layers have a large thermal resistance, they insulate the hot interior from the cold surrounding space,
throttling the energy loss L by photon radiation. Of course, WDs can
also lose energy by neutrino volume emission L , and by novel particle
emission Lx . Hence, WD cooling is governed by the equation
dU/dt = (L + L + Lx ).
(2.4)
This simple picture ignores the possibility of residual hydrogen burning near the surface, a possibly important luminosity source for young
WDs (e.g. Castellani, DeglInnocenti, and Romaniello 1994; Iben and
Tutukov 1984). I will get back to this problem below.
46
Chapter 2
(2.5)
The photon luminosity is L = 78.7 L 102Mbol /5 in terms of the bolometric magnitude, equivalent to log(L /L ) = (4.74 Mbol )/2.5. L is
related to the internal temperature T by the thermal conductance of
the surface layers so that one may derive a function T (L ). The quantities U , L , and Lx are given in terms of T so that they can be expressed
in terms of L and hence of Mbol .
In hot WDs the thermal energy is largely stored in the nuclei which
form a nearly classical Boltzmann gas. At low T the ideal-gas law
breaks down and eventually the nuclei arrange themselves in a lattice.
The internal energy is then a more complicated function of temperature.
The heat capacity per nucleon, which is 32 for an ideal gas, rises to
3 near the Debye temperature D and then drops approximately as
(16 4 /5) (T /D )3 to zero (Shapiro and Teukolsky 1983). However, the
observed WDs have a relatively small because of their small mass
around 0.6 M so that even the oldest WDs have not yet crystallized.
Therefore, as a reasonable rst approximation the internal energy is
U = C T with the ideal-gas heat capacity for the entire star of
C=
3 M Xj
L Gyr M Xj
= 3.95102
,
2 mu j Aj
107 K M j Aj
(2.6)
L = 1.710
M
L
M
T
107 K
)7/2
K T 7/2 ,
(2.7)
47
(2.8)
78.7L 102Mbol /5 + L + Lx
M
M
)5/7
Xj
j
Aj
(2.9)
M
M
)5/7
Xj
j
Aj
(2.10)
Taking M = 0.6 M and an equal mixture of
12
C and
16
O one nds
(2.11)
Neutrino Cooling
For the hottest WDs volume neutrino emission is more important than
surface photon cooling. The photon luminosity of Eq. (2.7) can be
expressed as an eective energy-loss rate per unit mass of the star, =
L /M = 3.3103 erg g1 s1 T73.5 with T7 = T /107 K. For the upper
relevant temperature range, neutrinos are emitted mostly by the plasma
48
Chapter 2
49
Fig. 2.11. Luminosity function for 0.6 M WDs for two values of 12 (Blinnikov 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 temperature bins of Tab. 2.2 as a function of the anomalous neutrino cooling implied
by a magnetic dipole moment (Blinnikov and Dunina-Barkovskaya 1994).
The number of WDs in the temperature range 12,00040,000 K is normalized to unity. The shaded bands correspond to the observations (Tab. 2.2).
50
Chapter 2
Standard or exotic neutrino emission from WDs (or neutron stars) has
the important property that it switches o quickly as the star cools
because of the steep temperature dependence of the emission rates.
Therefore, neutrinos cause a dip at the hot end of the luminosity function while older WDs are left unaected, even for signicantly enhanced
neutrino cooling (Fig. 2.11). One may construct other cases, however,
where this is dierent. One example is when a putative low-mass boson is emitted in place of a neutrino pair, say, in the bremsstrahlung
process e + (Z, A) (Z, A) + e + . The reduced nal-state phase
space then reduces the steepness of the temperature dependence of the
energy-loss rate. The possible existence of such particles is motivated
by theories involving spontaneously broken global symmetries. The
most widely discussed example is the axion which will be studied in
some detail in Chapter 14. For the present discussion all that matters
is the temperature variation of an assumed energy-loss rate.
The bremsstrahlung rate for pseudoscalar bosons will be calculated
in Chapter 3. For the highly degenerate limit the result is given in
Eq. (3.33) where = g 2 /4 is the relevant ne-structure constant.
Because this rate depends on the density only weakly through a factor
51
(2.12)
In the original derivation (Raelt 1986b) ion correlations were ignored, leading
to F 3 and to the limit 26 <
0.3.
52
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 26 where
I have corrected from F = 3, which they used in order to compare
with Raelt (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 ne-structure constant
26 = /1026 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 12,00040,000 K is normalized to
unity. The shaded bands correspond to the observations (Tab. 2.2).
53
P /P a T /T + b R/R
where the dimensionless constants a and b
are of order unity (Winget, Hansen, and van Horn 1983). Because
a WD has an almost xed radius, the second term may be ignored.
The time scale of cooling and of the period change are then related by
T /T = a P/P . ZZ Ceti stars have surface temperatures in the neighborhood of 13,000 K where the cooling time scale is of order 1 Gyr.
For a period of a few minutes one is talking of a period decrease
P = O(1014 s s1 ), not an easy quantity to measure.
After upper limits on P had been established over the years for a
number of cases, Kepler et al. (1991) succeeded at a measurement for
the DAV star G117B15A (Tab. 2.3) using the Whole Earth Telescope
which allows for nearly 24 h a day coverage of a given object.
A variety of model calculations give P = 25 1015 s s1 , somewhat smaller than the measured value, i.e. the star appears to cool
Surface temperature
Luminosity
Bolometric brightness
Mass
Pulsation period
Period change
Te
log(L/L )
Mbol
M
P
P
P/P
13,200 K
2.3
10.49 mag
(0.49 0.03) M
(215.197,387 0.000,001) s
(12 4) 1015 s s1
(0.57 0.17) Gyr
54
Chapter 2
2.3
2.3.1
Neutron Stars
Late-Time Cooling
Neutron stars are born when the degenerate iron core of an evolved
massive star becomes unstable and collapses to nuclear densities, an
implosion which is partly reected at the core bounce and leads to a
type II supernova explosion (Sect. 2.1.8). These events and the rst few
seconds of neutron star cooling are discussed more fully in Chapter 11.
For a few seconds the star emits most of its binding energy in the form
of MeV neutrinos which were observed from SN 1987A. Afterward, the
temperature at the neutrino sphere (the analogue of the photosphere in
ordinary stars) has dropped so much that the detectors are no longer
sensitive to the neutrino ux although the star continues to cool by
surface neutrino emission.
After 10100 yr the internal temperature has dropped to about
9
10 K 100 keV where the neutron star becomes entirely transparent to neutrinos and continues to cool by neutrino volume emission.
After about 105 yr it reaches an inner temperature of about 2108 K,
55
a point at which photon emission from the surface becomes the dominant 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 stiness (model FP), the dotted
line for a sti model (PS). All nonstandard eects were ignored such as
nucleon superuidity, a meson condensate, magnetic elds, and so forth.
Temperatures and luminosities are local, ignoring the gravitational redshift.
56
Chapter 2
2.3.2
X-Ray Observations
Age [yr]
(in 1995)
Dist.
[kpc]
Pulsarb
3C58
(814)
(2.6)
Crab
941
1.72
0531+21
(r, o, x, )
Remnant
RCW 103
1500 500
Te
[106 K]
2.2 0.2
< 1.6
2.15 0.15
< 1.2
E
R
1
3
detected
0.95 0.15
1.6 0.2
E
R
1
4
4.2
(r, x)
12,000
0.5
083345
(r, o, x, )
Monogem
Ring (?)
110,000
(0.5)
0656+14
(r, x)
0.90 0.04
(0.30 0.05)
340,000 0.150.4
Geminga
(x, )
0.52 0.10
540,000
105552
(r, x, )
0.65 0.15
0.75 0.06
X
R
7
8
Vela X
a Adapted
0.51.5
4. Ogelman,
Finley, and Zimmermann (1993).
b Pulsed
57
58
Chapter 2
59
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 direct URCA process anywhere in the star, and by the o switch from
superuidity. As the occurrence of these eects depends on ne 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 eect which would accelerate cooling is the occurrence of
a meson condensate (Sect. 4.9.1) because of the increased eciency of
neutrino emission. Again, this eect depends sensitively on the equation 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).
2.3.4
60
Chapter 2
12
limit of B <
310 B (Sect. 6.5.6), I interpret these results to mean
that a neutrino magnetic dipole moment leaves neutron-star cooling
unaectedone less nonstandard eect to worry about.
The rough agreement between the reference cooling curves and the
data points in Fig. 2.15, notably for the old pulsars, suggests that nonstandard eects cannot be much more ecient than standard neutrino
cooling unless all of the old isolated pulsar surface temperatures have
been incorrectly assignednot a likely scenario. Therefore, it is clear
that these and future observations of cooling neutron stars will be pivotal as laboratories to study novel phenomena such as the occurrence
of nonstandard phases of nuclear matter.
However, at the present time it is not clear if the emission of weakly
interacting particles such as axions could still have an interesting impact
on neutron star cooling. At the present time it appears easier to make
the reverse statement that such cooling eects likely are not important
in view of the restrictive limits on the interaction strength of axions or
nonstandard neutrinos set by other astrophysical objects. Therefore,
at present it appears that for the more narrowly dened purposes of
particle physics the role of old neutron stars as laboratories is less useful
than had been thought in the early works discussed above. It also
appears that novel weakly interacting particles usually would have a
more dramatic impact on the rst few seconds of Kelvin-Helmholtz
cooling of a protoneutron star (Chapter 11) than they do on the cooling
of old pulsars.
2.4
2.4.1
Globular-Cluster Stars
Observables in the Color-Magnitude Diagram
61
62
Chapter 2
fast. Therefore, the stars along the red-giant branch (RGB), horizontal 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 xed by helium 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. ConseTO
quently, the brightness dierence VHB
between the HB and the TO
is a distance-independent measure of the TO mass and thus of the cluster 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 lter (e.g. visual brightness V ) to a bolometric brightness, i.e. there is no need for a bolometric correction (BC).
Also, RR Lyrae stars are bright and easily identied 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
determinations of Buonanno, Corsi,
As an example the recent VHB
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
dened in Eq. (2.15). The best linear t is
TO
= (3.54 0.13) (0.008 0.078) [Fe/H],
VHB
(2.13)
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 species 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 xed by the treatment
of convection. Still, the TO color is a useful measure for the relative
ages between clusters of identical metallicities. Notably, the distanceindependent color dierence (B V ) between the TO and the base
of the RGB (Fig. 2.16) has been used to establish an age dierence 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 ignition. 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 brightness dierence between the RGB tip and the HB. However, because
the color of RR Lyrae stars and red giants is very dierent one must
64
Chapter 2
tip
convert to an absolute bolometric brightness dierence MHB
rather
tip
than using the visual brightness dierence VHB
.
Fig. 2.18. Evolutionary speed on the RGB for a model with M = 0.8 M ,
metallicity Z = 104 , and initial helium abundance Y = 0.240 (Raelt 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).
65
2.4.2
Theoretical Relations
In order to test the standard stellar-evolution picture against the observables introduced in Sect. 2.4.1 they need to be related to stellar
properties such as mass and chemical composition. In the past, extensive grids of stellar evolutionary sequences have been calculated and
have been used to derive analytic approximations for the connection
between various stellar parameters; as a canonical standard I use the
evolutionary HB and RG sequences of Sweigart and Gross (1976, 1978).
66
Chapter 2
a) Composition Parameters
The chemical composition is characterized by the helium content and
the metallicity. The convective envelope on the RGB at some point
reaches down into the region of a variable composition prole caused
by nuclear burning, causing a certain amount of processed material
(helium) to be dredged up. Therefore, the helium mass fraction Yenv
of the envelope is slightly larger than the value Y at formation. The
amount of dredge-up is approximately given by (Sweigart, Renzini, and
Tornambe 1987)
Ys Yenv Y 0.0136 + 0.0055 Z13 ,
(2.14)
In their calculation of RG sequences, Sweigart and Gross (1978) used the symbol
Y to denote the MS helium abundance; the envelope abundance near the helium
ash can be inferred from their tabulation of Ys values for each sequence. In
their 1976 study of HB sequences, they used the symbol Y to denote the envelope
abundance which is here consistently called Yenv .
67
(2.15)
(2.16)
(2.17)
where all stellar masses are understood in units of the solar mass M .
Here, Mc is the core mass at helium ignition and M7 M 0.7 is
the reduced total mass. I have increased Mc by 0.004 relative to
the original calculation to account for the corrected plasma neutrino
emission rate (Haft, Raelt, and Weiss 1994). Within about 0.003 M
Raelt and Weiss (1992) found the same expression (when corrected for
the plasma rates) except for a slightly shallower metallicity dependence.
Recently, Sweigart (1994) has reviewed the core-mass calculations
at the helium ash. All workers seem to agree within a few 103 M
except for Mazzitelli (1989) who found core masses larger by some
0.020 M . Sweigart (1994) claims that this disagreement cannot be
attributed to the algorithm adopted to accomplish the shell shifting
of the numerical grid which represents the star on a computer.
Because of substantial mass loss on the RGB the meaning of the
total mass M in this equation is not entirely obvious. If there were
enough time to relax to equilibrium one would think that it is the instantaneous mass at the helium ash. Indeed, Raelt and Weiss (1992)
found in an evolutionary sequence with mass loss that the end mass determined Mc . However, in this calculation the mass loss was stopped
68
Chapter 2
(2.19)
slightly dierent from the results of Raelt (1990b) who used the mass
of RR Lyrae stars for the total mass in Eq. (2.18).
Recall that the absolute bolometric brightness is given by M = 4.74 2.5 log L
for L in units of L , M in magnitudes.
11
69
(2.20)
(2.21)
70
Chapter 2
(at the RR Lyrae strip) and the RGB tip for which one nds with
Eqs. (2.19) and (2.20)
tip
MHB
MRR Mtip
(2.22)
(2.23)
Observational Results
71
However, because there are relatively few stars near the tip on the
RGB (on average about 10 mag1 in the observed clusters), the brightest RG is on average about 0.1 mag below the actual RGB tip. Raelt
(1990b) estimated the richness of the RGB near the tip for each cluster
on the basis of the rst few brightest stars provided by the observations and thus estimated the expected brightness dierence between
the brightest RG and the tip. This yields a linear regression
tip
= (4.19 0.03) + (0.41 0.06) Z13 ,
MHB
(2.25)
about 0.13 mag brighter than the t 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 modication.
More recent observations are those of Da Costa and Armandro
(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 coecient of the metallicity dependence agrees well
with the predicted value one may restrict a further comparison betip
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 nds
4.4 Y23 + RR 4.5 Mc = 0.06 0.03.
(2.26)
If RR = 0.2 mag this result implies that the envelope helium abun-
72
Chapter 2
(2.27)
Fig. 2.21. Number ratio of HB/RGB stars for 15 globular clusters according
to Buzzoni et al. (1983).
73
(2.28)
(2.29)
(2.30)
74
Chapter 2
Most recently, the brightness of RR Lyrae stars in the Large Magellanic Cloud was measured; it has a distance which is thought to be
well determined by other methods. Walker (1992) found
MRR = 0.48 + 0.15 Z13 ,
(2.31)
(2.32)
(2.33)
tip
from MHB
,
from R,
from MRR .
(2.34)
75
tip
from MHB
,
from R,
from MRR .
(2.35)
Bands of allowed values for Yenv and Mc are shown in Fig. 2.22.
From Fig. 2.22 one reads o that approximately
Yenv = 0.24 0.02
and
(2.36)
Fig. 2.22. Allowed values for the envelope helium abundance of evolved
globular-cluster stars and of an anomalous core-mass excess at helium ignitip
tion. The limits were derived from the observed brightness dierence 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 eld
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
(2.37)
An Alternate Analysis
The above discussion of the globular-cluster limits is a somewhat updated version of my own previous work (Raelt 1990b). Very recently,
Catelan, de Freitas Pacheco, and Horvath (1995) have provided an independent new and extended analysis. While they closely follow the
line of reasoning of Raelt (1990b) they have changed numerous detip
tails. Of the 26 globular clusters which enter the MHB
argument,
and of the 15 clusters which enter the R-method, they have discarded
several with an extreme HB morphology. For the absolute brightness
of RR Lyrae stars they have employed Walkers (1992) values which
depend on a precise knowledge of the LMC distance. For the brightness dierence between zero-age HB stars and RR Lyrae stars they use
RR = 0.31 + 0.10 [Fe/H] = 0.18 + 0.10 Z13 .
77
tip
from MHB
,
from R,
from MRR ,
from A.
(2.38)
Bands of allowed values for Yenv and Mc are shown in Fig. 2.23 which
is analogous to Fig. 2.22.
From this analysis one infers a best-t value for the envelope helium 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 eects
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 ignition according to the analysis of Catelan, de Freitas Pacheco, and Horvath
(1995). For comparison see Fig. 2.22.
78
Chapter 2
Systematic Uncertainties
The globular-cluster argument yields some of the most restrictive limits on novel modes of energy loss. Therefore, it is important to understand some of the systematic uncertainties that enter the nominal limit
Mc <
0.025 M .
Core rotation has often been quoted as an eect to change Mc .
However, as it would actually delay helium ignition it cannot be invoked
to compensate for anomalous cooling eects. In addition, if fast core
rotation were an important eect one would expect it to vary from star
tip
,
to star, causing a random broadening of the distribution of MHB
tip
is
an eect not indicated by the observations. The scatter of MHB
completely within observational errors and within the scatter caused
by the eect that the brightest red giant is not exactly at the tip of the
RGB in a given cluster. Further discussions of the rotational impact
on Mc are found in Catelan, de Freitas Pacheco, and Horvath (1995).
The uncertainty of the conductive opacities are relatively large as
stressed by Catelan, de Freitas Pacheco, and Horvath (1995). However,
conceivable modications of Mc do not seem to exceed the 0.010 M
level even with extreme assumptions.
One of the main theoretical weaknesses of the helium-ignition argument is that the helium ash has never been properly calculated.
Because helium ignites o-center one expects that convection plays a
major role in the process of heating the entire core and its expansion.
One may worry that in the process of the ash, parts of the core are
ejected into the stellar envelope, reducing its post-ash size. However,
if signicant amounts of helium were ejected, the inferred Yenv would be
changed dramatically. Even the ejection of 0.010 M of helium would
increase Yenv by 0.03 if one assumes an envelope mass of 0.3 M and
thus would brighten RR Lyrae stars by 0.12 mag. Within the stated
limit of Mc < 0.025 M , mass ejection from the core is a dramatic
eect that would be hard to hide in the data.
One signicant systematic uncertainty arises from the relative abundance of metals among each other which is usually xed by the solar
mixture (Ross and Aller 1976). The assumption of a Ross-Aller mixture for metal-poor systems like globular-cluster stars has been called
79
into question in recent years (e.g. Wheeler, Sneden, and Truran 1989).
It is thought that in these systems the elements (mostly oxygen)
are enhanced relative to iron. RG sequences calculated by Raelt and
Weiss (1992) with somewhat extreme enhancements indicated that
the red-giant core mass and luminosity at the helium ash are only
moderately changed. Most of the change that did occur was due to
the reduction of Fe at constant Z because of the enhancement. Put
another way, if Z13 = [Fe/H] + 1.3 is used as the dening equation for
the above reduced metallicity the eect of enhancements appear
to be rather minimal.
Catelan, de Freitas Pacheco, and Horvath (1995) have taken the
point of view that for enhanced elements one should rescale the
metallicity according to a recipe given by Chie, Straniero, and Salaris
(1991). Put another way, the metallicity parameter of Eq. (2.16) that
was used in the previous sections should be redened as
Z13 [Fe/H] + 1.3 + log(0.579 f + 0.421).
(2.39)
2.5
2.5.1
80
Chapter 2
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 consumption while the overall stellar structure remains nearly unchanged.
The temperature dependence of the helium-burning (triple- reaction)
energy generation rate 3 T 40 is much steeper than the case of hydrogen burning discussed in Sect. 1.3.1 and so the adjustment of the
stellar structure is even more negligible. With L3 the standard heliumburning 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 approximate factor L3 /(Lx + L3 ). Demanding a reduction by less than
10% translates into a requirement Lx <
0.1 L3 . Because this constraint is relatively tight one may compute both Lx and L3 from an
unperturbed model. If the same novel cooling mechanism has delayed
the helium ash 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 L3 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 3 80 erg g1 s1 . Then
a nonstandard energy-loss rate is constrained by
1 1
x <
10 erg g s .
(2.40)
Previously, this limit had been stated as 100 erg g1 s1 , overly conservative 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 = 104 g cm3 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 , and a somewhat larger power of the temperature
T . For the HB star model of Fig. 2.4 the core-averaged values n and
T n are shown in Fig. 2.24 as a function of n. The dependence on n
is relatively mild so that the nal result is not sensitive to ne points
of the averaging procedure.
In order to test the analytic criterion Eq. (2.40) in a concrete example consider axion losses by the Primako process. The energy-loss rate
will be derived in Sect. 5.2.1. It is found to be proportional to T 7 /
and to a coupling constant g10 = ga /1010 GeV1 . For a typical HB
81
Fig. 2.24. Average values of n and T n for the HB star model of Fig. 2.4
where 4 = /104 g cm3 and T8 = T /108 K.
2
core I nd T87 /4 0.3 which leads to g10
30 erg g1 s1 . Thus,
for g10 = 1 one concludes that the helium-burning lifetime should be
reduced by a factor 80/(80 + 30) = 0.7.
This result may be compared with numerical evolution sequences
for 1.3 M stars with an initial helium abundance of 25% and a metallicity of Z = 0.02 (Raelt and Dearborn 1987). The helium-burning
lifetime was found to be 1.2108 yr which was modied to 0.7108 yr
for g10 = 1. This means that the axion losses on the HB led to a tHe
reduction by a factor 0.6, in good agreement with the analytic estimate.
This comparison corroborates that the analytic criterion represents the
claimed impact of a novel energy loss on the helium-burning lifetime
with a reasonable precision.
Equation (2.40) may now be applied to many dierent cases. An
overview over the most salient results is given in Tab. 2.5 where the
original references are given, and the sections of this book are indicated where a more detailed discussion can be found. Apart from the
listed examples, the argument has also been applied to low-mass supersymmetric particles (Fukugita and Sakai 1982; Bouquet and Vayonakis
1982; Ellis and Olive 1983). However, it is now thought that supersymmetric particles, if they exist, are probably not light enough to be
produced in stars whence the interest in this case has waned.
82
Chapter 2
Particle property
Dominant process
Constrainta
[Ref.]
Sect.
Yukawa coupling of
Bremsstrahlung
gS < 1.31014
4
4
scalar (vector) boson e + He He + e + gV < 0.91014
to electrons
[1]
3.5.5
[1]
3.6.2
Yukawa coupling
to baryons
Compton
+ 4 He 4 He +
gS < 4.31011
gV < 3.01011
Photoproduction of
X boson
Photoproduction
+ 4 He 4 He + X
50
2
<
310 cm
Yukawa coupling of
pseudoscalar boson
a to electrons
Compton
+ee+a
g < 4.51013
[3]
3.2.6
Yukawa coupling of
paraphoton
Compton
+ e e +
g < 3.21013
[4]
3.2.6
coupling of
pseudoscalar boson
Primako
+ 4 He 4 He + a
ga < 0.61010
GeV1
[5]
5.2.5
Neutrino dipole
moment
Plasmon decay
e < 11011 B
[6]
6.5.6
a As
[2]
83
Helium Ignition
3 GN M 2
.
7 R
(2.41)
M
M
)1/3
M
.
M
(2.42)
From the numerical sequences of Sweigart and Gross (1978) one nds
0.81015 M s1
that near the helium ash M 0.5 M and M
so that grav 100 erg g1 s1 . Therefore, one must require x
100 erg g1 s1 in order to prevent the helium ash from being delayed.
In order to sharpen this criterion one may use results from Sweigart
and Gross (1978) and Raelt and Weiss (1992) who studied numerically
the delay of the helium ash by varying the standard neutrino losses
with a numerical factor F where F = 1 represents the standard case.
The results are shown in Fig. 2.25. Note that for F < 1 the standard
neutrino losses are decreased so that helium ignites earlier, causing
Mc < 0. It is also interesting that for F = 0 helium naturally
ignites at the center of the core while for F > 1 the ignition point
moves further and further toward the edge (Fig. 2.26). This behavior is
84
Chapter 2
85
quences the neutrino luminosity of the core at helium ignition is approximately 1 L so that 4 erg g1 s1 . Therefore, an approximate
analytic criterion to constrain a nonstandard energy loss is
1 1
x <
10 erg g s ,
(2.43)
[1026 ]
Mc [M ]
0.0
0.5
1.0
2.0
0.000
0.022
0.036
0.056
(2.44)
The same case was treated numerically by Raelt and Weiss (1995)
who implemented the energy-loss rate Eq. (3.33) with varying values
of in several red-giant evolutionary sequences.13 They found the
13
86
Chapter 2
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 (Raelt and Weiss
1995).
core-mass increases given in Tab. 2.6 and shown in Fig. 2.27. The requirement 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 available is the emission of neutrinos by the plasma process when
they have nonstandard magnetic dipole moments . The emission
rates are derived in Sect. 6.5.5 and the simple criterion Eq. (2.43) is
12
applied in Sect. 6.5.6. It yields a limit 12 <
2 where 12 = /10 B
with the Bohr magneton B = e/2me . The numerical variation of the
core mass with is shown in Fig. 2.28 according to Raelt and Weiss
(1992). An analytic approximation is
[
3/2
(2.45)
(2.46)
87
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 Raelt and
Weiss (1992) with a total stellar mass 0.8 M and an initial helium abundance of Y = 0.22 or 0.24 (the core-mass increase is found to be the same).
The triangles refer to the metallicity Z = 103 , the squares to 104 . The
open circles are the corresponding results of Castellani and DeglInnocenti
(1992) with the same stellar mass, Y = 0.23, and Z = 2104 . The solid
line is the analytic t Eq. (2.45).
2.6
Summary
88
Chapter 2
nonstandard energy-loss mechanism in stars are the white-dwarf luminosity function, the helium-burning lifetime of horizontal-branch stars,
and the nondelay of helium ignition in low-mass red giants as observed
by the brightness of the tip of the red-giant branch in globular clusters. It was possible to condense the latter two arguments into two
exceedingly simple criteria, namely that an anomalous energy-loss rate
in the cores of HB stars as well as in red-giant cores before helium
ignition must not exceed about 10 erg g1 s1 . The emission rate is to
be calculated at the pertinent plasma conditions, i.e. at an approximate temperature of 108 K = 8.6 keV, an electron concentration of
Ye = 0.5, and an average density of about 0.6104 g cm3 (HB stars)
or 2105 g cm3 (red giants). The former case corresponds to roughly
nondegenerate conditions, the latter case to degenerate ones so that it
depends on the density dependence of the emission rates which of these
cases will yield a more restrictive limit.
A red-giant core is essentially a 0.5 M helium white dwarf. The
observed real white dwarfs have typical masses of about 0.6 M ;
they are thought to consist mostly of carbon and oxygen. Both the
helium-ignition argument and the white-dwarf luminosity function allow one to constrain a novel energy-loss mechanism roughly on the level
of standard neutrino emission. Therefore, it is no surprise that bounds
derived from both arguments tend to be very similar.
There may be other objects or phenomena in the universe that measure novel particle-physics hypotheses even more sensitively than the
cases discussed here. They still need to make their way into the particle
astrophysics literature.
Chapter 3
Particles Interacting with
Electrons and Baryons
The stellar energy-loss argument is applied to weakly interacting particles which couple to electrons and baryons. The emission from a normal stellar plasma can proceed by a variety of reactions, for example
the Compton process e e where stands for a single particle
(axion, paraphoton, etc.) or a neutrino pair . Other examples of
practical interest are electron bremsstrahlung e (Z, A) (Z, A)e
or e e e e , electron free-bound transitions, and pair annihilation e e+ or e e+ . The corresponding energy-loss rates
of stellar plasmas are studied for temperatures and densities which are
of interest for stellar-evolution calculations. For particles coupled to
baryons some of the same processes apply in a normal plasma if one substitutes a proton or a helium nucleus for the electron. Reactions specic
to a nuclear medium are deferred to Chapter 4. In addition to draining
stars of energy, scalar or vector bosons would mediate long-range forces.
Leptonic long-range forces would be screened by the cosmic neutrino
background. Thermal graviton emission from stars is mentioned.
3.1
Introduction
90
Chapter 3
(Pontecorvo 1959; Gandelman and Pinaev 1959) or by photoproduction e e (Ritus 1961; Chiu and Stabler 1961). The set of
processes important for normal stars was completed by Adams, Ruderman, and Woo (1963) who discovered the plasma process .
Neutrino emission by these reactions is now a standard aspect of stellarevolution theory.
If other weakly interacting particles were to exist which couple directly to electrons they could essentially play the same role and thus
add to the energy loss of stars. The main speculation to be followed
up in this chapter is the possible existence of weakly interacting bosons
that would couple to electrons. Among standard particles the only lowmass bosons are photons and probably gravitons. The former dominate
the radiative energy transfer in stars, the latter are so weakly interacting that their thermal emission is negligible (Sect. 3.7). Why worry
about others?
Such a motivation arises from several sources. Low-mass bosons
could mediate long-range forces between electrically neutral bodies for
which gravity is the only standard interaction. It is an interesting
end in itself to set the best possible bounds on possible other forces
which might arise from the exchange of novel scalar or vector bosons.
Their existence seemed indicated for some time in the context of the
fth-force episode alluded to in Sect. 3.6.3 below. It is also possible
that baryon or lepton number play the role of physical charges similar
to the electric charge, and that a new gauge interaction is associated
with them. The baryonic or leptonic photons arising from this hypothesis are intriguing candidates for weakly interacting low-mass bosons
(Sect. 3.6.4). It will turn out, however, that typically massless bosons
which mediate long-range forces are best constrained by experiments
which test the equivalence principle of general relativity. Put another
way, to a high degree of accuracy gravity is found to be the only longrange interaction between neutral bodies.
Long-range leptonic forces can be screened by the cosmic neutrino
background. In this case the stellar energy-loss argument remains of
importance to limit their possible strength (Sect. 3.6.4).
The remaining category of interesting new bosons are those which
couple to the spin of fermions and thus do not mediate a long-range
force between unpolarized bodies. In the simplest case their CP-conserving coupling would be of a pseudoscalar nature. Such particles arise
naturally as Nambu-Goldstone bosons in scenarios where a global chiral
U(1) symmetry is spontaneously broken at some large energy scale. The
most widely discussed example is the Peccei-Quinn symmetry that was
91
3.2
3.2.1
Compton Process
Vector Bosons
92
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
s + 1 2 (
s2 6 s 3)
= 0
+
+
ln(
s) .
(3.1)
(
s 1)2
s2
(
s 1)3
with
g 2 /4,
(3.2)
(3.3)
1
s 1)/ s in the CM frame,
2 (
= 1
me
(
s 1)
in the electron rest frame.
2
93
photon eld tensor. In the nonrelativistic limit this yields the same total 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, polarization vector ) it is g f |eikr (k ) |i while for pseudoscalars it is
g f |eikr k |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 nal-state
vector, scalar, or pseudoscalar boson according to Eqs. (3.1), (3.5), and
(3.9), respectively, with 0 dened in Eq. (3.2) and the CM initial photon
energy.
3.2.2
Scalars
Grifols and Masso (1986) studied the stellar emission of scalars which
couple according to
Lint = g e e .
Integrating their dierential cross section I nd
[
= 0
(3.4)
]
1 3
s (
s + 3)2
16
+
+
ln(
s)
(
s 1)2
2
s2
(
s 1)3
(3.5)
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
3.2.3
Chapter 3
Pseudoscalars
(3.6)
Mpseudoscalar =
f e ki .
(3.7)
2me 2
This is to be compared with the corresponding matrix element for photon transitions,
Mphoton =
e 1 ikr
f e [2 p + (k )]i ,
2me 2
(3.8)
with the photon polarization vector and the electron momentum operator 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 rst 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 eective photon
mass relevant for a stellar plasma. For the present purpose it is enough
to consider massless photons and pseudoscalars. With 0 as dened in
Eq. (3.2) one nds
(
= 0
s1
ln(
s) 3
s 1
2
s2
(3.9)
(3.10)
95
Neutrino Pairs
(3.11)
= 0 (CV2 + CA2 )
+ (CV2 CA2 )
,
+ =
49 5 (13 s 7) 15 117 s 55 s3
+
+
12
(
s 1)2
12 s2
25 28 s 27 s2 2 s3 + 2 s4
+
ln(
s),
(
s 1)3
= 39 +
0 =
120 s
8 s 1 12 (2 + 2 s + 5 s2 + s3 )
+
ln(
s),
(
s 1)2
s2
(
s 1)3
G2F m2e
.
9 (4)2
(3.12)
6
35
(CV2 + 5CA2 ) (
s 1)4
= 0
96
35
(3.13)
55
24
55
48
],
(3.14)
2
With CV2 = CA
= 1 this result agrees with that of Ritus (1961) while Chiu and
Stabler (1961) appear to have an extra factor 2
s/(
s + 1). In the nonrelativistic limit
with s 1 this deviation makes no dierence while in the extreme relativistic limit
their result is a factor of 2 larger than that of Ritus (1961) and Dicus (1972).
96
Chapter 3
3.2.5
Energy-Loss Rates
The Compton-type processes are typically important when the electrons are nondegenerate (otherwise bremsstrahlung dominates) and
nonrelativistic (otherwise e+ e annihilation dominates). In these limits one may use the cross sections without Pauli blocking corrections.
Because the recoil of the target electron is neglected, the energy of
a photon impinging on an electron is identical with the energy carried
away by the new boson or neutrino pair. Therefore, the energy-loss
rate per unit volume is a simple integral over the initial-state photon
phase space, weighted with their Bose-Einstein occupation numbers,
Q = ne
2 d3 k
,
(2)3 e/T 1
(3.15)
97
initial- and nal-state electrons have the same momentum, reducing the
calculation to an average of the Pauli blocking factor over all electrons
Fdeg
(
)
1 2 d3 p
1
1
=
1 (E)/T
,
ne (2)3 e(E)/T + 1
e
+1
(3.16)
ex
1
p
E
dE
,
ne 2 me
(ex + 1)2
(3.17)
(3.18)
(3.19)
Here, n = (n) is the Riemann zeta function which shall be set equal
to unity.16 In a medium of mass density the electron density is ne =
Ye /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 T p+4
.
2
mu mpe
(3.20)
The average energy of the photons which are converted into weakly
interacting particles is
= (p + 3) T.
(3.21)
98
Chapter 3
8 Ye T 4
= 5.71029 erg g1 s1 Ye T84 ,
mu m2e
(3.22)
160 Ye T 6
= 3.31027 erg g1 s1 Ye T86 .
mu m4e
(3.23)
(CV2
5CA2 )
96 G2F m6e
T
Ye
4
mu
me
)8
(3.24)
Because p = 4 for this process, 7T . Relativistic corrections become important at rather low temperatures. For example, at T = 108 K
the true emission rate is about 25% smaller than given by Eq. (3.24).
3.2.6
99
<
0.91028
1.61026
scalar,
pseudoscalar.
(3.25)
These bounds apply to bosons with a mass below a few times the temperature, m <
2030 keV. For larger masses the limits are signicantly
degraded because only the high-energy tail of the blackbody photons
can produce the particles. For massive pseudoscalars this eect was
explicitly studied in Sect. 1.3.5 in the context of solar limits.
For vector bosons which interact by means of a Yukawa coupling,
the same limits to apply except that they are more restrictive by a
factor of 2 because of the two polarization states which increases the
emission rate.
For vector bosons which couple by means of a magnetic moment
as the paraphotons in Eq. (3.3), the bound on g is the same as for
pseudoscalars apart from an extra factor of 2 in the emission rate from
the two polarization states.
3.3
Pair Annihilation
2
s
=
ln
2
s 4me
4m2e
1+
1 4m2e /s
)]
(3.26)
.
12
1 4m2e /s
(3.27)
Because pair annihilation requires the presence of positrons it is important only for relativistic plasmas.
In this limit s1 ln(s) for the production of bosons (pseudoscalar, scalar, vector) and s for neutrino pairs. Therefore, the
100
Chapter 3
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.
importance of stellar energy-loss rates into new scalars relative to neutrino pairs decreases with increasing temperature and density. The
impact of new bosons on stellar evolution relative to neutrinos is then
expected to be most pronounced for low-mass stars where other processes such as photoproduction dominate. Consequently, the pair process has not played any signicant role at constraining the interactions
of new bosons.
3.4
101
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
3.5.1
Bremsstrahlung
Nondegenerate, Nonrelativistic Medium
In a typical stellar plasma there are many more charged particles than blackbody photons. In the solar center, for example, the electron density is 61025 cm3
while for photons at a temperature of 1.3 keV it is 2(3) T 3 / 2 = 61022 cm3 .
102
Chapter 3
128 2
Q=
45 me
[
nj Zj2 2
T
me
)5/2
ne
5 kS2
1
8 me T
+ Zj
5 kS2
1
4 me T
)]
(3.28)
Here, the sum is extended over all nuclear species with charges Zj e and
kS2
4
2
=
ne +
n j Zj ,
(3.29)
me T
me T 2
j
which is about 0.12 at the center of the Sun and 0.17 in the cores of
horizontal-branch (HB) stars.
Ignoring screening eects, the energy-loss rate per unit mass is
)
(
= 5.91022 erg g1 s1 T82.5 Ye
Yj Zj2 + Zj / 2 , (3.30)
j
8
103
Grifols, Masso, and Peris (1989) have worked out the bremsstrahlung rate for a scalar boson; in this case e e collisions can be ignored
relative to electron-nucleus scattering. They found in the nonrelativistic and nondegenerate limit, ignoring screening eects which are small,
=
22 ne T 1/2 Xj Zj2
a
3/2 mu m3/2
Aj
e
j
Xj Zj2
j
Aj
(3.31)
E1
42 2
d2 da
Q=
Zj n j
dE1 f1
dE2 (1 f2 )
2
4
4
me
me
j
[
2
|p1 | |p2 | 2
P1 K P2 K
2 P1 P2 me + (P2 P1 )K
2 2
2
+2
,
2
q (q + )
(P1 K)(P2 K)
P2 K P1 K
(3.32)
104
Chapter 3
where
F =
d2
4
(3.33)
da
(1 F2 ) [2 (1 c12 ) (c1a c2a )2 ]
4 (1 c1a F ) (1 c2a F ) (1 c12 )(1 c12 + 2 )
(3.34)
3
2
15
2
3 F
(3.36)
= 1.081027 erg g1 s1
ln
.
(3.38)
F2
2F3
1 F
The function f (F ) is 0 at F = 0 and rises monotonically to 1 for
105
2 + 2
2 + 2
ln
F =
2
2
1,
(3.39)
somewhat dierent from what Iwamoto (1984) found who used the
Thomas-Fermi wave number as a screening scale.
For a strongly coupled, degenerate plasma typical for white dwarfs
the factor F was calculated numerically by Nakagawa, Kohyama, and
Itoh (1987) and Nakagawa et al. (1988) who also gave analytic approximation formulae for the axion emission rate, applicable to nonrelativistic and relativistic conditions. For a 12 C plasma with densities in the
range 104 106 g cm3 and temperatures of 106 107 K it is found that
F = 1.0 within a few tens of percent. Therefore, for simple estimates
this value is a satisfactory approximation.
Altherr, Petitgirard, and del Ro Gaztelurrutia (1994) have calculated the bremsstrahlung process with the methods of nite temperature and density (FTD) eld theory. The main point is that one computes directly the interaction of the electrons with the electromagnetic
eld uctuations which are induced by the ambient charged particles.
The result for the emission rate is similar to the one derived above.
3.5.3
Neutrino pair bremsstrahlung (Fig. 3.5) was the rst nonnuclear neutrino emission process ever proposed (Pontecorvo 1959; Gandelman
and Pinaev 1959). A detailed calculation of the energy-loss rate in
a degenerate medium, relativistic and nonrelativistic, was performed
by Festa and Ruderman (1969) while conditions of partial degeneracy
were studied by Cazzola, de Zotti, and Saggion (1971). The Festa and
Ruderman calculation was extended by Dicus et al. (1976) to include
neutral-current interactions. After a calculation very similar to the one
presented above for pseudoscalars the emission rate for neutrino pairs is
(
)[
]
22 2 6
2
nj Zj 12 (CV2 + CA2 )F+ + 21 (CV2 CA2 )F .
Q=
GF T
189
j
(3.40)
With a single species of nuclei (charge Z, atomic mass A) the energyloss rate per unit mass is
= 0.144 erg g1 s1 (Z 2 /A) T86 [ . . . ],
8
(3.41)
where T8 = T /10 K and the square bracket is from Eq. (3.40). The
temperature dependence is steeper than for axions by two powers.
106
Chapter 3
The factors F+ and F are of order unity. Dicus et al. (1976) derived 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) result 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 eect is from the ion correlations which, in a weakly coupled plasma ( <
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 modication 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
eects) was pioneered by Flowers (1973, 1974) who showed clearly how
to separate the ion correlation eects 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 nd that F = 0 and
(
2 + 2
F+ ln
2
2
2 + 2
(3.42)
Neutron-Star Crust
107
One may now easily derive astrophysical limits on the Yukawa couplings
of scalars and pseudoscalars, in full analogy to Sect. 3.2.6 where the
Compton emission rates were used. I begin with the same case that was
1 1
considered there, namely the restriction <
10 erg g s in the cores
of horizontal-branch stars. For nondegenerate conditions the emission
rates are Eq. (3.30) and Eq. (3.31), respectively. They are proportional
to T 0.5 (pseudoscalar) and T 2.5 (scalar). With 4 = 0.64, T80.5 =
0.82, T82.5 = 0.48, and a chemical composition of pure helium one nds
<
1.41029
1.61025
scalar,
pseudoscalar.
(3.43)
Comparing these limits with those from the Compton process Eq. (3.25)
reveals that for scalars the present bremsstrahlung limit is more restrictive, for pseudoscalars the Compton one.
Because bremsstrahlung is not suppressed in a degenerate plasma,
one can also apply the helium-ignition argument of Sect. 2.5.2 which
1 1
8
again requires <
10 erg g s at the same T 10 K, however at a
density of around 106 g cm3 . For these conditions the plasma is degenerate but weakly coupled (Appendix D) so that Debye screening should
be an appropriate procedure. Therefore, for the emission of pseudoscalars the emission rate Eq. (3.35) with F from Eq. (3.36) should
be a reasonable approximation. The screening scale is dominated by
the ions, kS = ki = 222 keV, the Fermi momentum is pF = 409 keV so
that 2 = 0.15 while F = 0.77, yielding F 1.8.
In Fig. 3.6 the energy-loss rate of a helium plasma at T = 108 K
is plotted as a function of density, including the Compton process,
and the degenerate (D) and nondegenerate (ND) bremsstrahlung rates.
This gure claries that bremsstrahlung, of course, is suppressed by
degeneracy eects relative to the ND rates, but it is not a signicantly
decreasing function of density. In this gure a simple interpolation
(solid line) between the regimes of high and low degeneracy is shown
108
Chapter 3
that was used in a numerical study of axion emission from red giants
by Raelt 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
1 1
independent of density. The requirement <
10 erg g s then yields
26
a constraint <
for any density whence the helium-burning life 10
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 heliumignition argument at an average core density of 2105 g/cm3 (the central density is about 106 g/cm3 ) and at the almost constant temperature of T = 108 K. The degenerate emission rate is then approximately
21027 erg g1 s1 so that
26
<
0.510 .
(3.44)
109
The same case was treated numerically by Raelt and Weiss (1995)
who implemented the compound energy-loss rate of Fig. 3.6 with varying values of in several red-giant evolutionary sequences. The results of this work have been discussed in Sect. 2.5.2 where it was used
as one justication for the simple 10 erg g1 s1 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.
3.6
3.6.1
The main concern of this chapter was novel bosons which interact with
electrons by a dimensionless Yukawa coupling. It will become clear
shortly that these results can be easily translated into limits on baryonic couplings as well. It may be useful to pull together the main
results found so far, and discuss them in the context of other sources
of information on the same quantities.
The best studied case of boson couplings to electrons is that of
pseudoscalars because the existence of such particles is motivated by
their role as Nambu-Goldstone bosons of a spontaneously broken chiral
symmetry of the fundamental interactions. Within this class, axions
(Chapter 14) have been most widely discussed; they usually serve as a
generic example for low-mass pseudoscalars. For vector bosons which
couple by means of a magnetic moment (Eq. 3.3) such as paraphotons the same limits apply apart from an extra factor of 2 in the
emission rate from the two polarization states of these particles.
The simplest constraint on the Yukawa coupling ga of pseudoscalars
(axions) to electrons (a = ga2 /4) arises from the argument that the
age of the Sun precludes any novel energy-loss mechanism to be more
ecient than the surface photon luminosity (Sect. 1.3.2). The relevant
emission processes are the Compton reaction with the energy-loss rate
given in Eq. (3.23), and bremsstrahlung with electrons scattering on
electrons, protons, and helium nuclei; the emission rate was given in
Eq. (3.28). An integration over a typical solar model yields an axion
luminosity (Raelt 1986a)
La = a 6.01021 L ,
(3.45)
with about 25% from the Compton process, 25% from ee bremsstrahlung, and 50% from bremsstrahlung by electrons scattering on nuclei.
110
Chapter 3
22
<
The requirement La <
L yields a constraint a 1.710 .
A much more restrictive limit arises from the white-dwarf luminosity function as axion emission would accelerate white-dwarf cooling.
This argument was studied in detail in Sect. 2.2.4 as a generic case
for the use of the white-dwarf luminosity function; the resulting constraint is given in Tab. 3.1. The cooling speed of white dwarfs was
also established from a measurement of the period decrease of the DA
variable (ZZ Ceti) star G117B15A which thus yields a similar limit.
However, the period decrease of this star may be slightly faster than
can be attributed to standard cooling processes; it has been speculated
that axions with a coupling strength of about a 0.51026 could be
responsible (Sect. 2.2.5). The limit Eq. (3.44) that was derived in the
previous section from the helium-ignition argument in globular clusters
is of a similar magnitude, but not restrictive enough to exclude this
hypothesis.
All of these constraints apply to low-mass bosons. The most restrictive one is based on the helium ignition argument with T 108 K =
8.6 keV. Therefore, these constraints apply if m <
10 keV. However,
it would be incorrect to think that for larger masses there was no constraint. There is one, but it is degraded because threshold eects limit
the particle production to the high-energy tails of the thermal distributions of the plasma constituents. For massive pseudoscalars, this
question has been studied in Sect. 1.3.5 in the context of general solar
particle constraints. For the more restrictive globular-cluster limits,
such a detailed investigation does not exist in the literature.
3.6.2
The couplings of low-mass scalar or vector particles are easy to constrain by the same methods. Because the energy-loss rates have been
calculated only for nondegenerate conditions, only the arguments involving the solar age and the helium-burning lifetime can be employed.
The latter yields a constraint on the -e coupling of
29
e <
1.410 .
(3.46)
0.51026
1.61026
similar
Helium-burning lifetime of
horizontal-branch stars
Solar age
1.71022
1.01026
Astrophysical
observable
Upper limit
a = ga2 /4
Bremsstrahlung (degenerate)
e+AA+e+a
(A = 4 He, 12 C, 16 O)
Compton
+ee+a
same
Bremsstrahlung (degenerate)
e + 12 C (16 O) 12 C (16 O) + e + a
Dominant emission
process
Sect. 2.5.2
Sect. 3.5.5
Sect. 3.2.6
Sect. 2.2.5
Sect. 2.2.4
here
Detailed
discussion
Tab. 3.1. Astrophysical bounds on the Yukawa coupling ga of pseudoscalars (axions) to electrons.
112
Chapter 3
One may also consider the Yukawa coupling of scalar (vector) bosons
to baryons (Grifols and Masso 1986; Grifols, Masso, and Peris 1989b).
In this case one may use the Compton process + 4 He 4 He + 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 41 for pure
helium, me is to be replaced by mHe 4mu , and 4 to account
for the coherent photon coupling. Pulling these factors together one
nds = N 0.71023 erg g1 s1 for pure helium. Because the heliumburning lifetime argument limits this energy-loss rate to 10 erg g1 s1
one nds
22
N <
(3.47)
1.510
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.
3.6.3
Long-Range Forces
Scalar or vector particles mediate long-range forces between macroscopic bodies. For pseudoscalars this is not the case because their CPconserving coupling to fermions has a pseudoscalar structure, i.e. in the
nonrelativistic limit they couple to the fermion spin. Therefore, even
if the mass of the new particles is very small or exactly zero, they do
not mediate a long-range force between unpolarized bodies. The residual force caused by the simultaneous exchange of two pseudoscalars is
found to be extremely small (e.g. Grifols and Tortosa 1994).
Consider a scalar of mass m = 1 (Compton wave length ) which
couples to nucleons with a Yukawa strength g. It mediates an attractive force between two nucleons given in terms of the potential
(g 2 /4) r1 er/ . Two macroscopic test bodies of geometric dimension much below are then attracted by virtue of the total potential
m1 m2
(1 + er/ ),
(3.48)
V (r) = GN
r
19
where GN = m2
Pl is Newtons constant, mPl = 1.2210 GeV is the
Planck mass, m1 and m2 are the masses of the bodies, and
g 2 m2Pl
.
(3.49)
4 u1 u2
Here, u1,2 = m1,2 /N1,2 with N1,2 the total number of nucleons in each
body. Apart from small binding-energy eects which are dierent for
113
dierent materials u1 u2 mu is the atomic mass unit, approximately equal to a nucleon mass mN . For this force not to compete
with gravity one needs g < O(1019 ) or = g 2 /4 < O(1037 ) so that
low-mass scalars, if they exist, need to have extremely feeble couplings
to matter. Their smallness implies that such particles would not have
any impact whatsoever on the energy loss of stars.
Therefore, for boson masses so small that the Compton wave length
is macroscopic, the most restrictive limits on g obtain from analyzing
the forces between macroscopic bodies, not from the energy-loss argument. The existence of a composition-dependent fth force in nature
with a strength of about 1% of gravity and a range of a few hundred meters seemed indicated by a reanalysis of Eotvoss original data
(Fischbach et al. 1986). Subsequently many experiments were carried
out to search for this eect, with no believable positive outcome (for
a review see Fischbach and Talmadge 1992). However, these investigations produced extremely restrictive limits on , depending on the
assumed range of the new force. The limits also depend on the presumed coupling; if the new force couples to baryon number one nds
3
< 9
that <
10 for in the cm range, or 10 for of order the
Earth-Sun distance and above. Thus, novel long-range forces must be
much weaker than gravity. No bounds on seem to exist for below
the cm range, i.e. for boson masses of order 103 eV and above, apart
from the stellar energy-loss argument.
The eect of a novel force with intermediate range on the equations of stellar structure was discussed by Glass and Szamosi (1987,
1989). Solar models including such a force and the impact on the
solar oscillation frequencies were discussed by Gilliland and Dappen
(1987) and Kuhn (1988). For a force so weak or weaker than indicated by the laboratory limits, no observable consequences for stellar structure and evolution seem to obtain. Also, the impact of the
new eld on the value of fundamental coupling constants even at compact objects such as neutron stars is far below any observable limit
(Ellis et al. 1989).
A signicant bound obtains from the orbital decay of the HulseTaylor binary pulsar. In order for the energy loss in the new scalars to
remain below 1% of the gravitational wave emission the Yukawa cou19
pling to baryons must satisfy g <
(Mohanty and Panda 1994).
310
<
This translates into 1 for scalar boson masses below the orbital pul19
sar frequency of 2/P = 2.251104 s1 , i.e. for 1 <
1.510 eV
14
or >
1.310 cm. This fth-force limit, however, is weaker than
those derived by terrestrial laboratory methods.
114
3.6.4
Chapter 3
Leptonic and Baryonic Gauge Interactions
The physical motivation for considering long-range interactions mediated by vector bosons, besides the fth-force episode, is the hypothesis
that baryon number or lepton number could play the role of physical
charges similar to the electric one (Lee and Yang 1955; Okun 1969).
Their association with a gauge symmetry would provide one explanation for the strict conservation of baryon and lepton number which so
far has been observed in nature. In the framework of this hypothesis
one predicts the existence of baryonic or leptonic photons which couple
to baryons or leptons by a charge eB or eL , respectively. The novel
gauge bosons would be massless like the ordinary photon.
Therefore, the limits established in the previous sections on the dimensionless couplings of vector bosons can be readily restated as limits
on the values of putative baryonic or leptonic charges. The energy-loss
argument applied to helium-burning stars yields
14
eL <
110 ,
11
eB <
(3.50)
310 ,
according to Eqs. (3.43) and (3.47), respectively. Tests of the equivalence principle (i.e. of a composition-dependent fth force) on solar9
system scales yield <
10 (Sect. 3.6.3) so that
23
eB <
(3.51)
110 .
Apparently this is the most restrictive limit on eB that is currently
available.
One may be tempted to apply the limits from the equivalence principle 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 lled with a background neutrino sea in the same
way as it is lled with a background of microwave photons. This neutrino 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; C
iftci, Sultansoi,
and T
urkoz 1994; Dolgov and Raelt 1995).
Debye screening will be studied in Sect. 6.4.1. For the screening
wave number one nds the expression
dp fp p (v + v 1 ),
(3.52)
115
3
3
density is given by n+ = 2 fp d p/(2) = 0 fp p2 dp/ 2 . Therefore,
in the relativistic limit one nds roughly
1/3
1
kS1 e1
L n+ eL 0.2 cm,
(3.53)
116
Chapter 3
3.7
Chapter 4
Processes in a Nuclear
Medium
The interaction rates of neutrinos and axions with nucleons in a nuclear
medium are studied with a focus on neutral-current processes such as
bremsstahlung emission of axions N N N N a and of neutrino pairs
N N N N as well as neutrino scattering. A severe problem with
the perturbative rate calculations at high density is discussed which has
a strong impact on axion emissivities and the dominant axial-vector
contribution to the neutrino opacities.
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 p pa;
the previous results can be easily adapted to such reactions. Presently
I will focus on processes involving neutrinos or axions that are specic
to a nuclear medium, i.e. to supernova (SN) cores or neutron stars.
The main focus of the literature which deals with microscopic processes 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 identications 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 scenarios, notably those that involve novel phases of nuclear matter such
as superuidity, meson condensates, quark matter, etc. (Shapiro and
Teukolsky 1983; Tsuruta 1992).
117
118
Chapter 4
119
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.
4.2
4.2.1
16 (4)3 2 a
k2
|M| =
3m2N
k2 + m2
spins
2
)2
l2
+ 2
l + m2
)2
k2 l2 3 (k l)2
. (4.2)
+ 2
(k + m2 )(l2 + m2 )
Here, a (CN mN /fa )2 /4 and (f 2mN /m )2 /4 15 with f
1 are the axion-nucleon and pion-nucleon ne-structure constants,
respectively. Further, k = p2 p4 and l = p2 p3 with pi the momenta
of the nucleons Ni as in Fig. 4.1.
In a thermal medium k2 3mN T so that [k2 /(k2 + m2 )]2 varies
between 0.86 for T = 80 MeV and 0.37 for T = 10 MeV. Therefore,
neglecting the pion mass causes only a moderate error in a SN core
(Brinkmann and Turner 1988; Burrows, Ressell, and Turner 1990).
120
Chapter 4
Raelt and Seckel (1995) showed that including m causes less than
a 30% reduction of typical neutrino or axion rates for T > 20 MeV. Because this will be a minor error relative to the dominant uncertainties
l)2 ].
the term in square brackets is approximated as [3 (k
l)2 term is inconvenient without yielding any
The remaining (k
signicant 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 (Raelt 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 (4)3 2 a m2
N .
(4.3)
spins
While this may seem somewhat arbitrary, it must be stressed that using 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.
4.2.2
Energy-Loss Rate
The axionic volume energy-loss rate of a medium is the usual phasespace integral,
Qa =
4
d3 pi
d3 ka
f f (1 f3 )(1 f4 )
a
3 1 2
2a (2)3
i=1 2Ei (2)
(2)4 4 (P1 + P2 P3 P4 Ka ) 14
|M|2 , (4.4)
spins
where P1,2 are the four-momenta of the initial-state nucleons, P3,4 are
for the nal states, and Ka is for the axion. The factor 41 is a statistics
121
(4.5)
In this case, the second integral expression in Eq. (4.4), which knows
about axions only by virtue of the energy-momentum transfer Ka in
the function, is only a function of the axion energy a .
Therefore, in terms of a dimensionless function s(x) of the dimensionless axion energy x = a /T one may write the energy-loss rate in
the form (baryon density nB )
(
Qa =
=
CN
2fa
)2
nB
d3 ka
a s(a /T ) ea /T
2a (2)3
a n B T 3
dx x2 s(x) ex .
4 m2N
0
(4.6)
Nondegenerate Limit
3/2 p2 /2mN T
3
(nB /2) (2/mN T ) e
so that 2fp d p/(2)3 = nB gives the
nucleon (baryon) density where the factor 2 is for two spin states. Pauli
blocking factors are omitted: (1 f3,4 ) 1.
122
Chapter 4
5/2
= 4 2 nB T 1/2 mN ,
s(x) = 4
dy ey |x|y + y 2
)1/2
1 + |x| /4 .
(4.7)
Table 4.1. sn =
(D) conditions.
sn (ND)
sn (D)
1
2
3
4
8/5
23 + 65 / 2
128/35 31 4 /315
256/21 245 + (180/ 2 ) 7
4096/77 82 6 /315
123
explicitly
QND
=
a
,
9/2
35
mN
3.5
ND
= a 1.691035 erg g1 s1 15 TMeV
,
a
(4.8)
Degenerate Limit
Details of the nucleon phase space in the degenerate limit can be found
in Friman and Maxwells (1979) calculation of emission. The axion
energy-loss rate is expressed as in Eq. (4.6). One nds
=
42 pF T 3
3 nB
and
s(x) =
(x2 + 4 2 ) |x|
,
4 2 (1 e|x| )
(4.9)
124
Chapter 4
31 2 pF T 6
,
945 m2N
2/3
31
1 1
s 15
D
a = a 1.7410 erg g
6
,
TMeV
(4.10)
(4.11)
2
5u
u2
G(u) = 1
arctan
+
+
6
u
3(u2 + 4)
(
)
u2
2 2u2 + 4
+ 2
arctan
,
u2
6 2u + 4
(4.12)
Eq. (4.12) diers from the corresponding result of Friman and Maxwell (1979)
which is identical with that of Iwamoto (1984) who apparently did not take the
third term of the matrix element Eq. (4.2) properly into account.
125
44
153
T
mN
)4
(4.13)
For a mixed medium of protons and neutrons one needs to consider the
individual Yukawa couplings gan = Cn mN /fa and gap = Cp mN /fa as
well as the isoscalar and isovector combinations g0,1 = 21 (gan gap ) and
the ne-structure constants j = gj2 /4 with j = n, p, 0, 1. For equal
couplings n = p = 0 while 1 = 0.
In the nondegenerate limit, the main dierence is that np scattering
benets from
the exchange of charged pions which couple more strongly
by a factor 2. Depending on the chemical composition of the medium,
the emission rate will be increased by up to a factor of 2. On the other
l)2 term
hand, some reduction factors have been ignored such as the (k
and the pion mass. Therefore, ignoring this enhancement essentially
compensates for the previously introduced errors.
In the degenerate limit the changes are more dramatic. The role
of the Fermi momentum is played by pF (3 2 nB )1/3 . It sets the
126
Chapter 4
scale for the proton and neutron Fermi momenta which are pn,p
=
F
1/3
(3 2 nB )1/3 Yn,p
. According to a result of Brinkmann and Turner (1988)
the eective coupling is then
+
a n Yn1/3 + p Yp1/3 + ( 28
3 0
20
) (Yn2/3
3 1
1
2 2
+ Yp2/3 )1/2
|Yn2/3 Yp2/3 |
2/3
Yn
2/3
+ Yp
(4.14)
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.
4.3
4.3.1
Neutrino pair emission in nucleon-nucleon collisions (Fig. 4.5) is analogous to that of axions. Therefore, instead of embarking on a new
calculation it is worth understanding their common featuresthe neutrino rates can be obtained for free on the basis of the axion ones.
This approach amounts to dening the dynamical structure functions
127
Fig. 4.5. Bremsstrahlung emission of neutrino pairs in nucleon-nucleon collisions. There is a total of eight amplitudes, four with the neutrinos attached
to each nucleon line, and an exchange graph each with N3 N4 .
of the medium, quantities which are of utmost importance to understand the general properties of the emission, absorption, and scattering
rates independently of phase-space details of the neutrinos or axions.
Contrary to the bremsstrahlung emission in electron-nucleus collisions (Sect. 3.5.3), nonrelativistically only the nucleon axial-vector
coupling contributes in nucleon-nucleon collision (Friman and Maxwell
1979). This dierence originates from the interaction potential of the
colliding particles which involves a spin-dependent force between nucleons so that the spin uctuations caused by collisions are more dramatic
than those of the velocitysee Sect. 4.6.5 below. Hence, the part of the
interaction Hamiltonian relevant for neutrino pair bremsstrahlung is
Hint =
CAN GF
N 5 N (1 5 ) ,
2
(4.15)
where CAp CAn 21 ; see Appendix B for a discussion of the appropriate values in a nuclear medium.
The interaction Hamiltonian Eq. (4.15) has the same structure as
that for axions Eq. (4.1). The squared matrix elements are then of the
general form
A
(4.16)
|M|2 =
(C /2f )2 M K K
for axions.
spins
N
a
a a
Here, Ka is the axion four-momentum while
(
N = 8 K1 K2 + K2 K1 K1 K2 g i K1 K2
(4.17)
128
Chapter 4
4
1
d3 pi
f1 f2 (1 f3 )(1 f4 )
nB i=1 2Ei (2)3
(2)4 4 (P1 + P2 P3 P4 + K) M .
(4.18)
Here, the Pi are the nucleon four momenta, K = Ka for axion emission, and K = (K1 + K2 ) for neutrino pairs. Thus, K is the energymomentum transfer from the radiation (axions or neutrino pairs) to
the nucleons. Because S knows about the radiation only through
the energy-momentum function it is only a function of K = (, k),
apart from the temperature and chemical potentials of the medium.
The slightly awkward denition of the sign of K follows the common
denition of the structure function where a positive energy transfer
refers to energy given to the medium.
The energy-loss rates by or axion emission are then the phasespace integrals
(
Q =
(
Qa =
CAN GF
2
CN
2fa
)2
nB
)2
nB
d3 k1
d3 k2
S N (1 + 2 ),
21 (2)3 22 (2)3
d3 ka
S Ka Ka a .
2a (2)3
(4.19)
Here, it was assumed that both axions and neutrinos can escape freely
from the medium so that nal-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 i where is a nucleon two-spinor and i (i =
1, 2, 3) are Pauli matrices representing the nucleon spin operator. Put
another way, in the nonrelativistic limit the axial-vector current represents the nucleon spin density. Therefore, it has only spatial components so that S S ij (i, j = 1, 2, 3). In order to construct the
most general tensorial structure for S ij in an isotropic medium only ij
is available. Recall that in the nonrelativistic limit S does not know
about the momentum transfer k because of Eq. (4.5). There is then no
129
(4.20)
Q =
(
Qa =
CAN GF
2
CN
2fa
)2
)2
nB
d 6 S (),
20 4 0
nB
d 4 S ().
4 2 0
(4.21)
S () = 2 s(/T )
1
e/T
for > 0,
for < 0.
(4.22)
130
4.3.2
Chapter 4
Bremsstrahlung Emission of Neutrino Pairs
2 11/2
2048
2 2 2 nB T
C
G
,
385 7/2 A F m5/2
N
5.5
ND
= 2.41017 erg g1 s1 15 TMeV
,
(4.23)
ND
with 15 = /1015 g cm3 , TMeV = T /MeV, and ND
= Q / is the
energy-loss rate per unit mass.
For the degenerate rate one uses Eqs. (4.9). The average energy of
a neutrino pair is (s4 /s3 ) T 5.78 T and the total energy-loss rate is21
QD
=
41 2 2 2
CA GF pF T 8 ,
4725
2/3
13
1 1
D
s 15
= 4.410 erg g
8
TMeV
.
(4.24)
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 coupling is isovector (CAp CAn ) so that 0 0 while the other s are
approximately equal (Appendix B). For degenerate conditions, with a
small modication the dependence on the proton concentration is the
same as that shown in Fig. 4.4. Again, the absolutely dominating contribution is from np collisions unless protons are so rare that they are
nondegenerate.22
21
Friman and Maxwells (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
conict with the results of Friman and Maxwell (1979). They found that Q was
proportional to the proton Fermi momentum which is relatively small in neutronstar matter. On the other hand, for small proton concentrations Brinkmann and
Turners pF in Eq. (4.24) approaches the neutron Fermi momentum. I am in no
position to decide between these conicting results.
4.4
131
Axion Opacity
(4.25)
where is the mass density of the medium, the boson mean free path
against absorption, and B (T ) = (2 2 )1 3 (e/T 1)1 is the boson
spectral density for one spin degree of freedom. After T = /T has
been taken one nds
1
15
x4 e2x
= 4
,
(4.26)
dx x x
a
8 0
(e 1)3
where x /T , and (1 ex )1 = ex (ex 1)1 was used.
The axion opacity thus dened is to be added to the photon opacity
1
1
by 1
tot = + a . The energy ux is then given by the usual expression F = (3tot )1 a T 4 where a = 2 /15 gives the radiation
density of photons. Burrows, Ressell, and Turner (1990) have dened
an axion opacity which is to be used in conjunction with the axion radiation density which is half as large because there is only one spin degree
of freedom, i.e. aa = 2 /30. Their 1
a is twice that in Eq. (4.26) so
that the energy ux is the same. I prefer the present denition because
it allows for the usual addition of all opacity contributions.
It is easy to determine the axion absorption rate from the discussion
in Sect. 4.3.1. Starting
from Qa in Eq. (4.21) one removes the phase
space integral 0 d 4 2 /(2)3 and a factor because Qa was an
energy-loss rate. One includes a factor e/T to account for the detailedbalance relationship (see Eq. 4.43) so that altogether
(
CN
2fa
)2
nB s(x)
,
T
2x
(4.27)
a =
CN
2fa
)2
,
T mN
(4.28)
132
Chapter 4
15
8 4
dx
x4 e2x
2x
.
(ex 1)3 s(x)
(4.29)
For the nondegenerate case s(x) was given in Eq. (4.7), for the degenerate one in Eq. (4.9). Then one nds numerically
ND = 0.46 and
4.5
4.5.1
Neutrino Opacity
Elastic Scattering
(4.30)
Pair Absorption
133
=
=
CA GF
)2
nB
d3 k2
f2 S N
3
21 22 (2)
3CA2 G2F nB
d2 22 f2 S (1 + 2 ),
2 2
0
(4.31)
where 1 refers to the for which the mfp is being determined while
2 refers to a from the thermal environment (occupation number
f2 ). With the detailed-balance relationship S () = S () e/T (see
Eq. 4.43) and writing S (||) = ( / 2 ) s(/T ) as before one nds
3CA2
G2F
nB T
dx2 f2
x22 s(x1 + x2 )
,
2 2 (x1 + x2 )2
(4.32)
(4.33)
nB
16
= 4 1/2 2 5/2
T
0
mN T 1/2
30 MeV
T
)1/2
(4.34)
Inelastic Scattering
It appears that for typical conditions of a young SN core, pair absorption is almost as important as elastic scattering. However, even though
the quantity is larger than unity, the pair-absorption rate has an
unfavorable phase-space factor from the initial-state so that the dimensionless integral in Eq. (4.33) is a small number. This would not
be the case for the inelastic scattering process N N N N shown in
134
Chapter 4
Fig. 4.6. In this process, neutrinos can give or take energy even though
recoil eects for heavy nucleons are small in the elastic scattering process N N .
This reaction is identical with pair absorption with the antineutrino
line crossed into the nal state. The corresponding mfp is given by the
same expression as for pair absorption, except that the index 2 now
refers to the nal-state , the initial-state occupation number f2 is to
be replaced with a nal-state Pauli blocking factor (1 f2 ), and the
energy transfer is = 2 1 ,
3CA2 G2F nB
=
d2 22 (1 f2 ) S (1 2 ).
2 2
0
(4.35)
135
However, what exactly does one mean by a vanishing energy transfer? The nucleons in the ambient medium constantly scatter with each
other so that their individual energies are uncertain to within about
1/coll with a typical time between collision coll . Therefore, one would
expect that the structure of S (), which was calculated on the basis
of free nucleons which interact only once, is smeared out over scales
1
of order coll
. In particular, this smearing-out eect naturally
regulates the low- behavior of the structure function (Raelt and
Seckel 1991).
1
Because coll
is a typical nucleon-nucleon collision rate, a simple ansatz for a modied S () is a Lorentzian (Raelt and Seckel 1991)
S (||)
s(/T
)
=
s(x),
2 + 2 /4
T x2 + 2 /4
(4.36)
+
with x = /T . For 1 one has
d S () = 2 + O( ) for the
modied S (). Thus, for 1 essentially S () = 2() so that
the total inelastic scattering rate Eq. (4.35) reproduces the elastic
scattering one. At the same time S () has wings which, for ,
give the correct inelastic scattering rate which is of order .
This simple ansatz can be expected to give a reasonable approximation to the true S () only in the limit 1. For practical
applications in cooling calculations of young SN cores, however, one
has to confront the opposite limit 1 causing the smearing-out effect by multiple nucleon collisions to be a dominating feature of S ()
even for T . This implies that for typical thermal energies of
neutrinos and nucleons a reasonably clean separation between elastic
and inelastic scattering processes is not logically possiblethere is only
one structure function S (), broadly smeared out, which governs all
axial-vector scattering, emission, and absorption processes.
This observation has important ramications not only for neutrino
scattering, but also for the bremsstrahlung emission of axions and neutrino pairs from a nuclear medium. Earlier it seemed that one did not
have to worry about details of the behavior of S () near = 0 because the low- part was suppressed by axion or neutrino phase-space
factors. However, since the notion of low energy presently means
<
, and because T , all relevant energy transfers are low in
this sense, and even the emission processes are dominated by multiplescattering eects. Consequences of this behavior are explored in more
detail below after a formal introduction of the structure functions and
their general properties.
136
4.6
4.6.1
Chapter 4
Structure Functions
Formal Denition
To treat axion and neutrino pair emission and absorption on the same
footing it became useful in Sect. 4.3.1 to dene the quantity S which
was the nuclear part of the squared matrix element of nucleon-nucleon
collisions, integrated over the nucleon phase space. The structure function thus obtained embodied the medium properties relevant for dierent processes without involving the radiation phase space. Clearly, this
method is not limited to the bremsstrahlung process. For example, one
could include the interaction of nucleons with thermal pions or a pion
condensate, three-nucleon collisions, and so forth. Whatever the details of the medium physics, in the end one will arrive at some function
S (, k) of the energy and momentum transfer which embodies all of
its properties. High-density properties of the medium should also appear in the structure function so that multiple-scattering modications
can be consistently applied to all relevant processes such as neutrino
scattering, pair emission, axion emission, and others.
A formal denition of the structure function without reference to
specic processes begins with a neutral-current interaction Hamiltonian
Hint = (gV V + gA A ) J ,
(4.37)
1 +
dt eit V (t, k)V (0, k) ,
nB
(4.38)
137
d
1
SV (, k) =
V (k)V (k) .
2
nB
(4.39)
+
It was used that
d eit = 2(t) so that dt in Eq. (4.38) is
trivially done and yields the operators at equal times. In Eq. (4.39)
V (k) is V (t, k) at an arbitrary time, for example t = 0. It only matters
that both V (k) and V (k) are taken at equal times.
In an isotropic medium the tensorial composition of the dynamical
structure function can be obtained only from the energy-momentum
transfer K and the four-velocity U of the medium; U = (1, 0, 0, 0) in its
rest frame. The general form of the vector term is (Kirzhnits, Losyakov,
and Chechin 1990)
SV = S1,V U U + S2,V (U U g )
+ S3,V K K + S4,V (K U + U K ).
(4.40)
(4.41)
t fk =
CA
2fa
)2
]
nB [
(fk + 1) SA (, k) fk SA (, k) K K .
2
(4.42)
(4.43)
138
4.6.2
Chapter 4
Nonrelativistic Limit
(4.44)
1 +
dt eit (t, k)(0, k) ,
nB
S (, k) =
4 +
dt eit s(t, k) s(0, k)
3nB
(4.45)
(4.46)
(4.47)
i=j
139
For the static spin-density structure function the same steps can be
performed with the inclusion of the spin operators 12 i for the individual
nucleons. This takes us to the equivalent of Eq. (4.47)
|s(k) s(k)| =
NB
1
ekrij 14 | i j | =
V i,j=1
NB
NB
1
1
2
1
|
|
+
eikrij 14 | i j |.
i
V i=1 4
V i,j=1
(4.48)
i=j
(4.49)
The structure functions have a number of general properties, independently of details of the interactions of the medium constituents. We
have already seen that they must obey the normalization condition
d
1
S (, k) = 1 +
2
nB
d
4
S (, k) = 1 +
2
3nB
NB
cos(k rij ) ,
i,j=1
i=j
NB
i j cos(k rij ) .
(4.50)
i,j=1
i=j
140
Chapter 4
then yields
+
d
d +
S (, k) =
(4.51)
d
1 [
(k), H (k) ,
S (, k) =
2
nB
]
d
4 [
S (, k) =
(k), H (k) .
2
3nB
(4.52)
Here it was used, again, that d eit = 2(t) and (k) (0, k) and
(k) (0, k).
In order to evaluate this sum rule more explicitly one must assume
a specic 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
1
V (rij , i , j ),
H=
+
2 i,j=1
i=1 2mN
NB
p2i
(4.53)
i=j
where again rij ri rj . One can then proceed to evaluate the commutators in Eq. (4.52). By virtue of the continuity equation for the particle
number one can then show (Pines and Nozi`eres 1966; Sigl 1995b)
d
k2
S (, k) =
,
2
2mN
d
k2
S (, k) =
2
2mN
4
+
3nB
NB [
i,j=1
i=j
141
For the spin-density structure function one can go one step further
by expressing the most general nonrelativistic interaction potential as
(Sigl 1995b)
V (rij , i , j ) = U0 (rij ) US (rij ) i j
[
(4.55)
d
k2
S (, k) =
2
2mN
4
+
3nB
NB [
1 cos(k rij )
VijS
+ 1 + cos(k rij )
1
2
VijT
. (4.56)
i,j=1
i=j
Long-Wavelength Properties
In calculations involving the emission or scattering of neutrinos or axions as in Sect. 4.3.1 one usually neglects the momentum transfer k
because the medium constituents are so heavy that recoil eects are
142
Chapter 4
(4.57)
k0
d
d
4
S () =
(1 + e/T )S () = 1 +
2
2
3nB
0
d
d
4
S () =
(1 e/T )S () =
2
2
3nB
0
NB
i j ,
i,j=1
i=j
NB
3 T
V
2 ij
i,j=1
i=j
(4.58)
These relations will be of great use to develop a general understanding
of the behavior of S () at high densities. The second column of expressions follows from the rst by detailed balance. Because S () 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 (t, k) and s(t, k) are
constant so that S, () = 2(), 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 property because in the long-wavelength limit it depends on (t, k0) =
V 1 d3 r (t, r) which remains constant. Therefore, even in an interacting medium one expects S () = 2(), in agreement with the
nding that the neutrino vector current does not contribute to bremsstrahlung in the nonrelativistic limit relative to the axial-vector current
(Friman and Maxwell 1979).
B
The relevant quantity for the latter is V 1 d3 r s(t, r) = 21 N
i=1 i
with i the individual nucleon spins. If the evolution of dierent spins
143
is uncorrelated one may ignore the cross terms in the correlator. With
the single-nucleon spin operator one nds then
S () =
1
3
(4.59)
CN
2fa
)2
4
12 2
(4.60)
for the dierential axion energy-loss rate (radiation power) per nucleon.
Because of collisions with other nucleons, and because of a spin dependent interaction potential caused by pion exchange, the nucleon spins
evolve nontrivially so that the correlator has nonvanishing power at
= 0, allowing for axion emission.
4.6.5
CN
2fa
)2
2
12 2
it
dt e
2
,
(t)
(4.61)
2
dI
2 +
it
,
=
dt
e
a(t)
d
3
(4.62)
144
Chapter 4
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
v = a. The role of k (axions) is played by the polarization
vector (photons) so that k2 = 2 must be replaced by 2 = 1. With
(CA /2fa ) e, using = e2 /4, and inserting a factor of 2 for two
photon polarization states completes the translation.
In order to understand the radiation spectrum consider a single
CN
2fa
)2
2
||2 .
2
12
(4.63)
(t)
=
i (t ti ).
(4.64)
i=1
CN
2fa
)2
2
coll ()2 F (),
12 2
(4.65)
where coll is the average collision rate and ()2 is the average
squared change of the spin in a collision. Further,
n
i j cos[(ti tj )]
1
,
F () = 1 + lim
n n
()2
i,j=1
(4.66)
i=j
where the rst term (the diagonal part of the double sum) gives the
total radiation power as an incoherent sum of individual collisions. The
145
it
is
suppressed.
The
low- suppression of brems coll
strahlung is known as the Landau-Pomeranchuk-Migdal eect (Landau
and Pomeranchuk 1953a,b; Migdal 1956; Knoll and Voskresensky 1995
and references therein).
The summation in Eq. (4.66) can be viewed as an integration over
the relative time coordinate t = ti tj with a certain distribution
function f (t). For a random sequence of kicks one expects an exponential distribution of the normalized form f (t) = 41 e t/2
where is some inverse time-scale. This implies the Lorentzian shape
(e.g. Knoll and Voskresensky 1995)
F () =
2
.
2 + 2 /4
(4.67)
A Lorentzian model is familiar, for example, from the collisional broadening of spectral lines. A comparison with Eq. (4.60) indicates that
coll ()2
=
2 + 2 /4
(4.68)
()2
coll .
2
(4.69)
146
4.6.6
Chapter 4
Classical vs. Quantum Result
In order to make contact with the quantum calculation that led to the
axion emission rate in Eq. (4.6) it is useful to juxtapose it with the
classical result in terms of the spin-density structure function. The
classical calculation led to
2
S () = 2 2
+ 2 /4
(4.70)
S () = 2 s(/T )
1
e/T
for > 0,
for < 0.
(4.71)
The function s(x) has the property s(0) = 1, it is even, and according
to the f-sum rule must decrease for large x. In Fig. 4.7 the classical and
quantum results are shown as dotted and dashed lines, respectively,
where for the purpose of illustration s(x) = (1 + x2 /4)1/4 has been
assumed.
Fig. 4.7. Classical, quantum, and compound spin-density structure function in the nondegenerate limit according to Eq. (4.70) and Eq. (4.71) with
/T = 0.2 and s(x) = (1 + x2 /4)1/4 .
This comparison highlights an important weakness of the classical result: it does not obey the detailed-balance requirement S () =
S () e/T . The classical correlators are invariant under time reection t t and thus symmetric under . Again, the classical
result is adequate only for || <
T.
147
Equation (4.71) also highlights an important weakness of the quantum result: It does not include the interference eect from multiple
collisions at || <
because the calculation was done assuming individual, isolated collisions. The normalization condition Eq. (4.58) can
then be satised only by accepting a pathological infrared behavior of
S () like the one suggested by Sawyer (1995).
The classical and the quantum results thus both violate fundamental
requirements. The quantum calculation applies for || >
while the
<
classical one for || T . If T (dilute medium) the regimes of
<
validity overlap for <
|| T and the results agree beautifully. In
this case the two calculations mutually conrm and complement each
other. The compound structure function, where the quantum result is
multiplied with 2 /( 2 +2 /4), fullls the detailed-balance requirement
and approximately the normalization condition. Therefore, it probably
is a good rst guess for the overall shape of S ().
4.6.7
High-Density Behavior
S (||) = 2
,
(4.72)
+ 2 /4
where is to be determined by the normalization condition. In the
dilute limit ( T ) this implies while in the dense limit
( T ) one nds /2.
148
Chapter 4
4 e/T
2 + 2 /4
(4.73)
Fig. 4.8. Schematic variation of the axion emission rate per nucleon with
/T , taking Eq. (4.72) for the spin-density structure function. The dashed
line is the naive rate without the inclusion of multiple-scattering eects, i.e.
it is based on S (||) = / 2 .
149
and may even decrease at large densities, although such large values for
may never be reached in a nuclear medium as will become clear below.
The high-density downturn of the axion emission rate can be interpreted in terms of the Landau-Pomeranchuk-Migdal eect (Landau
and Pomeranchuk 1953a,b; Feinberg and Pomeranchuk 1956; Migdal
1956) as pointed out by Raelt and Seckel (1991). The main idea is
that collisions interrupt the radiation process. The formation of a radiation quantum of frequency takes about a time 1 according to the
uncertainty principle and so if collisions are more frequent than this
time, the radiation process is suppressed. Classically, this eect was
demonstrated in the language of current correlators in Sect. 4.6.5.
The impact of the high-density behavior of S () on the neutralcurrent 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 MaxwellBoltzmann expression ei /T ; the resulting error is small for nondegenerate neutrinos. The relevant quantity is then
d1
d2 12 22 e1 /T S (1 2 ).
(4.74)
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 nds
(T 2 + T /2 + 2 /12) e/T
1
d
.
(4.75)
2 + 2 /4
0
This expression is constant for T where = and thus the
structure function is essentially 2(). Indeed, the average scattering
cross section (or mean free path) is not expected to depend on the
density.
For dense media ( T ), however, the broadening of S () beyond a delta function leads to a decreasing average scattering rate. This
means that at a xed temperature the medium becomes more transparent to neutrinos with increasing density, even without the impact
of degeneracy eects. 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 ipped in a collision with other nucleons because the interaction potential couples to the spin. Neutrino scattering with an energy transfer 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 /T , taking Eq. (4.72) for the spindensity structure function. The dashed line is the naive rate without the
inclusion of multiple-scattering eects, i.e. it is based on S () = 2().
This statement is based on the assumption that there are no spatial correlations
among the nucleonspossibly 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
specic nucleon interaction model showed a substantial suppression of S (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.
151
4.7
152
Chapter 4
(1989) argued on the basis of the nonlinear sigma model that the combination of parameters 2 a /m2N should remain approximately constant.
Mayle et al. (1989) similarly found that this parameter should remain
somewhere in the range 0.31.5 of its vacuum value.
The nucleon eective mass mN deviates substantially from the vacuum value mN = 939 MeV. This shift has an impact on the kinematics
of reactions and the nucleon phase-space distribution. A calculation of
mN has to rely on an eective theory which describes the interaction
of nucleons and mesons. A typical result from a self-consistent relativistic Brueckner calculation including vacuum uctuations is shown
in Fig. 4.10. For conditions relevant for a SN core, an eective value as
low as mN /mN = 0.5 is conceivable.
Fig. 4.10. Eective nucleon mass in a nuclear medium according to a selfconsistent relativistic Brueckner calculation (Horowitz and Serot 1987). Nuclear density corresponds to about 0 = 31014 g/cm3 .
4.8
Neutral-current weak processes are of prime interest for the topics studied in this book because they are closely related to exotic reactions
involving new particles such as axions. For completeness, however,
it must be mentioned that the charged-current reactions depicted in
Fig. 4.11 dominate the neutrino energy-loss rate of old neutron stars,
and also dominate the e opacity in young SN cores. The processes
n p e e (neutron decay) and e p n e are usually called the URCA
153
Fig. 4.11. The URCA processes. For the modied versions, there are obvious
other graphs with the leptons attached to other nucleon lines, and exchange
amplitudes.
457 2
GF cos2 C (1 + 3CA2 ) m2N pF,e T 6 ,
10080
(4.76)
154
Chapter 4
If the triangle condition is not satised the URCA processes require bystander particles to absorb momentum, leading to the modied URCA process shown in Fig. 4.11 (Chiu and Salpeter 1964). Of
course, as below in Sect. 4.9.1, the missing momentum can be provided
by pions or a pion condensate (Bahcall and Wolf 1965a,b).
An explicit result for the modied URCA rate in the OPE model
for the nucleon interactions is (Friman and Maxwell 1979)
Qmod.
URCA
11513 2 2
G cos2 C CA2 pF,e T 8 ,
120960 F
(4.77)
4.9
4.9.1
155
Npk
Np+k
+
,
Ep+k Ep Epk Ep
(4.78)
156
Chapter 4
where Np is a plane-wave nucleon state with energy Ep and spin orientation relative to k , and 0 is a coupling constant.
In the nonrelativistic limit the axion energy-loss rate corresponding
f N
f a is written as
to the decay N
1
2
Qa =
d3 k
d3 p1 d3 p2
f1 (1 f2 )
2(2)3
(2)3 (2)3
2 (E1 E2 )
|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
|M|2 =
spins
C0 geA C1
2fa
)2
4
3
A2 20
(2)3 3 (p + k ) + 3 (p k ) ,
(4.80)
Qa =
45
C0 geA C1
2fa
)2
A2 20 m2N T 4
,
|k |
(4.81)
2 20 T 4
= a A2 1.031044 erg cm3 s1 T94 ,
45 |k |
(4.82)
157
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 Tatsumi (1988) who found
Q =
2 e 2 2 2 2 20 T 6
C G m A
.
945 A F N
|k |
(4.83)
Quark Matter
158
Chapter 4
The gluon in Fig. 4.13 couples to the quarks with the strong nestructure constant s and the gluon propagator involves an eective
mass for which Anand, Goyal, and Iha (1990) used m2g = (6s /) p2F
where pF is the Fermi momentum of the quark sea. The axion coupling to quarks is the usual derivative form (Cq /2fa ) q 5 q with
q = u, d, s. After a cumbersome but straightforward calculation Anand
et al. found for the energy-loss rate in the degenerate limit
62 2 s2 pF T 6
Qa =
945 (2fa )2
(4.84)
s5 (1 + c4 )
1 2
d
d
,
2 2 0
(s2 s2 + 2 )2
0
I2
1 2
s5
d
d
,
2 2 0
(s2 s2 + 2 )(s2 c2 + 2 )
0
(4.85)
(4.86)
where the function G(u) was given in Eq. (4.12). The ratio of coupling
constants is model dependent (Sect. 14.3.3), the Fermi momenta are
approximately equal for equal densities.
For a typical value s = 0.4 one has s2 /2 = 0.6103 while the last
factor is about 140 with G = 0.7 (Fig. 4.3) so that the last two factors
together are about 0.08. Hence, the emission rate from quark matter
is much smaller than that from neutron matter because is so large
relative to s . In the inevitable presence of protons in a neutron star,
the np process will be even more important than nn (Fig. 4.4), further
enhancing the emission rate of nuclear matter relative to quark matter.
For neutrino pairs the ratio of the emission rates is about the same
as for axions and so quark matter is much less eective at emission
159
than nuclear matter (Burrows 1979; Anand, Goyal, and Iha 1990). This
observation has little eect 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 axion 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 coupling which allows for the gluonic Primako eect (Fig. 4.14) which is
analogous to the photon Primako eect discussed in Sect. 5.2.
Bubble Phase
160
Chapter 4
4.10
So far in this chapter neutrinos were assumed to be massless particles which interact only by the standard left-handed weak current. It
is possible, however, that neutrinos have a Dirac mass in which case
any reaction with a nal-state (anti)neutrino produces both positive
and negative helicity states. Typically, the wrong-helicity states will
emerge in a fraction (m /2E )2 of all cases because of the mismatch
between chirality (eigenstates of 5 ) and helicity. For E m the
wrong-helicity states correspond approximately to right-handed chirality states and so their interaction-rate with the ambient medium is
weaker by an approximate factor (m /2E )2 . This implies that for a
suciently small mass they would not be trapped in a SN core and
thus carry away energy in an invisible channel, allowing one to set
constraints on a Dirac neutrino mass from the observed neutrino signal
of SN 1987A (Sect. 13.8.1). Here, the relationship between the production rate of left- and right-handed neutrinos is explored in some
detail because the simple scaling with (m /2E )2 is not correct in
all cases.
When a neutrino interacts with a medium the transition probability
from a state with four-momentum K1 = (1 , k1 ) to one with K2 =
(2 , k2 ) is written as W (K1 , K2 ). The function W is dened for both
positive and negative energies. The emission probability for a pair
(K1 )(K2 ) is then W (K1 , K2 ), the absorption probability for a pair
is W (K1 , K2 ). The collisional rate of change of the occupation number
fk1 of a neutrino eld mode k1 is then given by
dfk1
=
dt coll
d3 k2 [
WK2 ,K1 fk2 (1 fk1 ) WK1 ,K2 fk1 (1 fk2 )
(2)3
]
161
G2F nB
S N ,
8 21 22
(4.88)
(4.89)
G2F nB [
(1 + cos )S1 + (3 cos )S2
4
(4.90)
i
i
i
(4.91)
for right-handed ones. After this substitution has been performed all
further eects of m are of higher order so that one may neglect m
everywhere except in the global spin-ip factor.
162
Chapter 4
m
2R
)2 [
163
Qscat =
Qpair =
d3 kL d3 kR f
WKL ,KR fkL R ,
(2)3 (2)3
d3 kL d3 kR f
WKL ,KR (1 f kL )R .
(2)3 (2)3
(4.93)
In the notation of Sect. 4.6 and for a single species of nucleons one has in this
2
limit S1 = CV2 S and S2 = CA
S .
26
2
In the notation of Sect. 4.6 this is S = CV2 S + 3CA
S in a medium consisting
of a single species of nucleons.
164
Chapter 4
[( )
]
G2F m2 nB T 4
2
+ 6 + 12 S(),
=
d
4 4
T
T
0
( )
G2F m2 nB T 4
1 4
=
d
S(),
4 4
12 T
0
(4.94)
27
Chapter 5
Two-Photon Coupling of
Low-Mass Bosons
Well-known particles such as neutral pions, or hypothetical ones such
as axions or gravitons, each have a two-photon interaction vertex. It
allows for radiative decays of the form 2 as well as for the
Primako conversion in the presence of external electric or
magnetic elds. The Primako conversion of photons into gravitons in
putative cosmic magnetic elds could cause temperature uctuations
of the cosmic microwave background radiation. In stars, the Primako
conversion of photons leads to the production of gravitons and axions,
although the graviton luminosity is always negligible. The Primako
conversion of axions into photons serves as the basis for a detection
scheme for galactic axions, and was used in several laboratory axion
search experiments. The modied Maxwell equations in the presence
of very low-mass pseudoscalars would substantially modify pulsar electrodynamics. These physical phenomena are explored and current laboratory and astrophysical limits on the axion-photon coupling are reviewed.
5.1
The idea that symmetries of the fundamental interactions can be broken by the vacuum or ground state of the elds plays an important
role in modern particle physics. The breakdown of a global symmetry by the vacuum expectation value of a Higgs-like eld leads to the
existence of massless particles, the Nambu-Goldstone bosons of the
broken symmetry. One well-known example are the pions which are
165
166
Chapter 5
(5.1)
(5.2)
167
the nucleons, for axions they are the quarks, possibly the charged leptons, 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 automatically 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)
ga =
=
,
m
f
(5.3)
where m was taken to be much larger than the axion and photon energies. 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.
For massive pseudoscalars the two-photon coupling allows for a decay a 2 with a width
2
a2 = ga
m3a /64.
(5.4)
For pions in the sigma model, the only charged fermion is the proton.
Then g = /f and 2 = 2 m3 /64 3 f2 = 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 eld
theory literature, e.g. Itzykson and Zuber 1983).
168
Chapter 5
The main point for the present discussion is that pseudoscalar massless or low-mass bosons are a natural consequence of certain extensions
of the standard model, and that these particles couple to photons according to Eq. (5.1) with a strength
ga =
Ca ,
fa
(5.5)
where fa is the energy scale of symmetry breaking and Ca is a modeldependent 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.
5.2
5.2.1
(5.6)
169
This cross section exhibits the usual forward divergence from the
long-range Coulomb interaction. For ma = 0 it is cut o in vacuum by
the minimum necessary momentum transfer qmin = m2a /2 (ma );
2 1
the total cross section is then a = Z 2 ga
[ 2 ln(2/ma ) 14 ].
In a plasma, the long-range Coulomb potential is cut o by screening
eects; according to Sect. 6.4 the dierential cross section is modied
with a factor q2 /(kS2 + q2 ). In a nondegenerate medium the screening
scale is given by the Debye-H
uckel formula
(
)
4
nB Ye +
Zj2 Yj ,
T
j
kS2 =
(5.7)
g2 T k2
= a S
32
[(
k2
1 + S2
4
4 2
ln 1 + 2
kS
1 ,
(5.8)
where the plasma mass of the initial-state photon and the axion mass
were neglected relative to the energy . In the limit kS this ex2
pression expands as a = ga
2 T /16 which is entirely independent
of the density and chemical composition.
For a stellar plasma, however, this approximation is usually not
justied. Ignoring the plasma frequency for the initial-state photons,
the energy-loss rate per unit volume is
Q=
2
ga
T7
2 d3 k a
=
F (2 ),
(2)3 e/T 1
4
(5.9)
dx (x2 + 2 ) ln 1 + 2 x2 x
, (5.10)
F (2 ) = 2
2 0
e 1
170
Chapter 5
5.2.2
T a =
d3 ka
1
d3 kL
ZT ZL |M|2 (2)4
3
3
2L (2) 2a (2) 2T
[
4 (KT + KL Ka ) 4 (KT KL Ka )
+
, (5.11)
eL /T 1
1 eL /T
where ZT,L are the vertex renormalization factors. Here, the rst term
in square brackets corresponds to coalescence and so it involves a BoseEinstein occupation number for the initial-state L . The second term
171
(5.12)
2
ga
a L
da da ZT ZeL
|kL kT |2
=
2
16 (2)
T
[
(T + L a ) (T L a )
+
.
eL /T 1
1 eL /T
(5.13)
2
T
ga
L kT |2 .
=
da ZT ZeL |k
2
8 (2)
(5.14)
In this limit a has the same energy as T while L has only provided
momentum.
To nish up, note that in the nondegenerate, nonrelativistic limit
ZT = ZeL = 1. Because P T and T = O(T ) we have P T
L kT |2 =
so that kT T . Moreover, kL = kT ka yielding |k
|ka kT |2 /|ka kT |2 = T2 (1 z 2 )/2(1 z) = 21 T2 (1 + z) where z is the
172
Chapter 5
2
ga
T
.
16
(5.15)
These simple calculations of the axion emission rate apply only in the
classical (nondegenerate, nonrelativistic) limit. Even though this is the
most relevant case from a practical perspective it is worth mentioning
how one proceeds for a more general evaluation. To this end note that
the 2 interaction Eq. (5.1) corresponds to a source term for the axion
wave equation,
( + m2a ) a = ga E B,
(5.16)
where = = t2 2 . Axions are then emitted by the E B uctuations caused by the presence of thermal electromagnetic radiation
as well as the collective and random motion of charged particles.
The Primako calculation in Sect. 5.2.1 used the (screened) electric
eld of charged particles and the magnetic eld of (transverse) electromagnetic radiation (photons) as a source. Actually, one can include
the magnetic eld of moving charges for this purpose. Then axions are
emitted in the collision of two particles (Fig. 5.4), a process sometimes
referred to as the electro Primako eect. Unsurprisingly, the emission rate is much smaller because the magnetic eld associated with
173
ga
ga
f () =
( k) d3 x eiqx E(x) =
( k) E(q),
4
4
(5.17)
where k and are the wave and polarization vector of the incident
wave, respectively, q is the momentum transfer to the axion, and
E(q) is a Fourier component of E(x). The dierential transition rate
is d/d = |f ()|2 so that
g2
da
= a 2 ( k)i ( k)j Ei (q)Ej (q),
d
(4)
(5.18)
where it was used that E(x) is real so that E (q) = E(q). If E(x) is
a random eld conguration one needs to take an ensemble average so
that the transition rate is proportional to Ei Ej q Ei (q)Ej (q).
The electric and magnetic eld uctuations of a plasma are intimately related to the medium response functions to electric and magnetic elds, i.e. to the polarization tensor. For a plasma at temperature T one can show on general grounds (Sitenko 1967)
Ei Ej q =
2
d
/T
2 e
1
(
qi qj Im L
qi qj
+ ij 2
2
2
q |L |
q
Im T
,
|T q2 / 2 |2
(5.19)
where L,T (, q) are the longitudinal and transverse dielectric permittivities of the medium (Sect. 6.3.3). A quantity such as Im L /|L |2 is
known as a spectral densityhere of the longitudinal uctuations.
174
Chapter 5
iq
j T /(1 + q2 /kS2 ) in the classiOne nds explicitly Ei Ej q = q
2
cal limit (Sitenko 1967) where kS is the Debye-H
uckel wave number of
Eq. (5.7). With this result one easily reproduces the Primako transition rate a (Raelt 1988a).
The language of spectral densities for the electromagnetic eld uctuations forms the starting point for a quantum calculation of the axion
emission rate in the framework of thermal eld theory. This program
was carried out in a series of papers by Altherr (1990, 1991), Altherr and
Kraemmer (1992), and Altherr, Petitgirard, and del Ro Gaztelurrutia
(1994). Naturally, in the classical limit they reproduced the Primako
transition rate T a of Eq. (5.8).
In the degenerate or relativistic limit their results cannot be represented in terms of simple analytic formulae. The most important
astrophysical environment to be used for extracting bounds on ga are
low-mass stars before and after helium ignition with a core temperature of about 108 K (Sect. 5.2.5). Altherr, Petitgirard, and del Ro
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-
175
(5.20)
with L the solar luminosity and g10 ga 1010 GeV. (Recalling that
ga = (/fa ) Ca this corresponds to fa /Ca = 2.3107 GeV.) The
dierential ux at Earth is well approximated by the formula
dFa
(a /keV)3
2
= g10
4.021010 cm2 s1 keV1 a /1.08 keV
da
e
1
(5.21)
which is shown in Fig. 5.6. The average axion energy is a = 4.2 keV.
2
The total ux at Earth is Fa = g10
3.541011 cm2 s1 .
The standard Sun is about halfway through its main-sequence
evolution. Therefore, the solar axion luminosity must not exceed its
Fig. 5.6. Axion ux at Earth according to Eq. (5.25) from the Primako
conversion of photons in the Sun.
176
Chapter 5
photon luminosity; otherwise its nuclear fuel would have been spent
before reaching an age of 4.5109 yr. This requirement yields a bound
1
9
ga <
2.410 GeV .
(5.22)
Globular-Cluster Bound on ga
Armed with the Primako emission rate Eq. (5.9) it is an easy task
to derive a bound on ga from the energy-loss argument applied to
globular-cluster stars (Sect. 2.5). We need to require that at T 108 K
the axionic energy-loss rate is below 10 erg g1 s1 for a density of about
0.6104 g cm3 , corresponding to a classical plasma, and for about
2105 g cm3 , corresponding to degeneracy. From Fig. 5.5 it is evident that the emission rate is a steeply falling function of density when
degeneracy eects become important. Obviously, the more restrictive
limit is found from the low-density case which is based on the heliumburning lifetime of HB stars (Sect. 2.5.1).
In order to calculate the average energy-loss rate of the core of an
HB star one needs T 7 / if in Eq. (5.9) one uses a constant 2 =
2.5 or F = 1.0. For a typical HB-star model (Fig. 1.4) one nds
T87 /4 0.3 where T8 = T /108 K and 4 = /104 g cm3 . There2
fore, a g10
30 erg g1 s1 so that the criterion Eq. (2.40) yields a
constraint
1
10
ga <
0.610 GeV
or
7
fa /Ca >
410 GeV.
(5.23)
5.3
One of the most interesting ramications of the electromagnetic coupling of pseudoscalars is the possibility to search for dark-matter axions.
It is briey explained in Chapter 14 that axions would be produced in
the early universe by a nonthermal mechanism which excites classical
177
178
Chapter 5
Fig. 5.7. Results of the galactic axion search experiments of the RochesterBrookhaven-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
51025 g cm3 = 300 MeV cm3 . The axion line is the relationship between axion mass and coupling strength for = 1 or E/N = 8/3 according
to Eq. (14.24).
5.4
5.4.1
179
Axion-Photon Oscillations
Mixing Equations
(5.24)
a
a because a term ga Ferad
may approximate ga Fe a ga Feext
is of second order in the weak radiation elds. Moreover, if only an external magnetic eld is present, the wave equation for the time-varying
part of the vector potential A and for the axion eld are
A = ga BT t a,
( m2a ) a = ga BT t A,
(5.25)
180
Chapter 5
n 1
nR
2
2
2
n 1
+ z + 2 nR
0
ga BT /2
A
0
ga BT /2 A = 0,
m2a /2 2
a
(5.26)
where the o-diagonal terms were made real by a suitable global transformation of the elds. Further, a photon index of refraction was included because in practice one never has a perfect vacuum.
The refractive index is generally dierent for the two linear polarization states parallel and perpendicular to BT (Cotton-Mouton eect).
Also, there may be mixing between the A and A elds, i.e. the plane
of polarization may rotate in optically active media, an eect characterized by nR . In general, any medium becomes optically active if there
is a magnetic eld component along the direction of propagation (Faraday eect). Therefore, in general the refractive indices n, depend on
the transverse, the index nR on the longitudinal magnetic eld.
Equation (5.26) is made linear by an approach that will be discussed
in more detail for neutrinos in Sect. 8.2. For propagation in the positive
z-direction and for very relativistic axions and photons one may expand
( 2 + z2 ) = ( + iz )( iz ) 2 ( iz ). Then one obtains the
usual Schrodinger equation
+ R
0
0
A
a + iz A = 0,
a
a
(5.27)
(5.29)
181
By adjusting the gas pressure within the magnetic eld volume one
can make the photon and axion degenerate and thus enhance the transition rate (van Bibber et al. 1989). This applies, in particular, to
solar axions which have keV energies so that the corresponding photon dispersion relation in low-Z gases is particle-like with the plasma
frequency being the eective mass.
If there is a gradient of the gas density, for example near a star, or
if the gas density and magnetic eld strength change in time as in the
expanding universe, suitable conditions allow for resonant axion-photon
conversions in the spirit of the neutrino MSW eect (Yoshimura 1988;
Yanagida and Yoshimura 1988).
For the magnetic conversion of pseudoscalars in the galactic magnetic eld one must worry about density uctuations 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 signicantly suppressed relative to the naive result.
5.4.2
Solar Axions
182
Chapter 5
However, an ongoing experimental project at the Institute for Nuclear Physics in Novosibirsk may be able to improve the helioscope
signicantly. The conversion magnet has been gimballed so that it can
track the Sun, providing much longer exposure times. First results can
be expected for late 1995see Vorobyov and Kolokolov (1995) for a
status report.
Another possibility would be to use the straight sections of the
beam pipe of the LEP accelerator at CERN as an axion helioscope.
Hoogeveen and Stuart (1992) have calculated the times and dates of
alignment with the Sun. They proposed an experimental setup that
might allow one to reach a sensitivity in ga down to 41010 GeV1 ,
which would be very impressive, but still far from the globular-cluster
bound Eq. (5.23).
Finally, Paschos and Zioutas (1994) proposed to use a single crystal
as a detector where the Primako conversion of solar axions is coherently enhanced over the electric elds of many atoms. Put another way,
one would expect a strong enhancement via Bragg scattering. Even
with this improvement, however, it does not seem possible to beat the
bound from globular-cluster stars.
5.4.3
Instead of using the solar axion ux one can make ones own by shining
a laser beam through a long transverse magnetic eld 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 eld (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 eld
strength of 3.7 T. The light beam was trapped in a resonant cavity
in the rst magnet, allowing for about 200 traversals; the incident
laser power was 1.5 W. At the end of the second magnet photons were
3
searched for by a photomultiplier. For an axion mass ma <
10 eV an
upper bound ga < 0.7106 GeV1 was found.
183
Vacuum Birefringence
]
22 [ 2
2 2
2
(E
B
)
+
7(E
B)
45m4e
(5.30)
(Heisenberg and Euler 1936; see also Itzykson and Zuber 1983).
Fig. 5.8. Vacuum birefringence in the presence of external elds. (a) QED
contribution according to the Euler-Heisenberg interaction. (b) a or
oscillations in an external E or B eld. (c) Photon birefringence in an
external axion eld (axionic domain walls, cosmic axion eld). (d) Axionmediated contribution in a strong E B eld, e.g. near a pulsar.
184
Chapter 5
22 B 2
45 m4e
and
n = 1 + 4
22 B 2
,
45 m4e
(5.31)
where
(22 /45) B 2 /m4e = 1.321032 (B/Gauss)2 .
(5.32)
The photon-splitting box graph with one external eld and three real photons
attached to an electron loop does not contribute. The lowest-order amplitude is
with the external eld attached three times, and three real photons (hexagon diagram). For references to the early literature and a discussion of the astrophysical
implications of the photon-splitting process see Baring (1991).
29
, refer to the electric eld of the wave relative to the external transverse B
eld while Adler (1971) refers with , to the magnetic eld of the wave. Note
also that I use rationalized units where = e2 /4 = 1/137 while in the literature
on photon refraction unrationalized units with = e2 = 1/137 are often employed.
185
4
a more restrictive limit of ga < 3.6107 GeV1 for ma <
710 eV.
For a larger mass an a- oscillation pattern develops on the length scale
of the optical cavity, leading to an oscillating limit as a function of
ma . In this regime the ellipticity measurement was superior.
New experimental eorts in the birefringence category include a
proposal by Cooper and Stedman (1995) to use ring lasers. A laser
experiment which is actually in the process of being built is PVLAS
(Bakalov et al. 1994) which will be able to improve previous laboratory
limits on ga by a factor of 40, i.e. it is expected to be sensitive in
1
8
< 3
the regime ga >
110 GeV as long as ma 10 eV. While such
strong couplings are astrophysically excluded it is intriguing that this
experiment should be able to detect for the rst time the standard QED
birefringence eect of Fig. 5.8a.
5.5
5.5.1
Certain stars have very strong magnetic elds. For example, neutron
stars frequently have elds of 1012 1013 G (e.g. Meszaros 1992), and
even white dwarfs can have elds of up to 109 G. Therefore, one may
think that axions produced in the hot interior of neutron stars at a temperature of, say, 50 keV would convert to -rays in the magnetosphere
(Morris 1986). However, the vacuum refractive term suppresses the
conversion rate because the photon momentum for a given frequency
is k = n, > with the refractive indices Eq. (5.31) while for the
axions ka = m2a /2 < . Therefore, in the presence of a magnetic
eld axions and photons are less degenerate so that it is more dicult
for them to oscillate into each other.
In principle, the refractive index can be cancelled by the presence
of a plasma where the photon forward scattering on electrons induces a
negative n1, i.e. something like a photon eective mass. In the aligned
rotator model for a magnetized neutron star a self-consistent solution of
the Maxwell equations with currents requires the presence of an electron
density of about ne = 71010 cm3 B12 Ps1 where B12 is the magnetic
eld along the rotation axis in units of 1012 G and Ps is the pulsar period
in seconds (Goldreich and Julian 1969). The corresponding plasma
frequency is P2 = 4ne /me = 0.971010 eV2 B12 Ps1 . This implies
1
with keV = /keV, to
k = P2 /2 = 51014 eV B12 Ps1 keV
2
keV which is
be compared with = (n 1) = 0.92104 eV B12
much larger, allowing one to ignore the plasma term.
186
Chapter 5
5.5.2
In the previous section it was shown that near a pulsar the QED vacuum Cotton-Mouton eect (Fig. 5.7a) induces a sizeable amount of
birefringence between the photon states which are linearly polarized
parallel or perpendicular to the transverse component of the magnetic
eld. Recently, Mohanty and Nayak (1993) showed that in addition
187
(5.33)
It appears as a source for the arion eld on the r.h.s. of Eq. (5.24). Taking account of the relativistic space-time metric outside of the pulsar,
Mohanty and Nayak (1993) found for the resulting arion eld
a = ga
(5.34)
where GN is Newtons constant and M the pulsar mass. The entire magnetosphere contributes coherently to this result. If the pseudoscalars had a mass, only the density E B within a distance of about
188
Chapter 5
m1
a would eectively act as a source for the local a eld and so it would
be much smaller.
An inhomogeneous pseudoscalar eld conguration represents an
optically active medium (Fig. 5.8c) as was noted, for example, in the
context of axionic domain wall congurations (Sikivie 1984). To lowest order the dispersion relation for left- and right-handed circularly
polarized light is (e.g. Harari and Sikivie 1992)
a,
k = 1 ga k
(5.35)
2
189
2
19
one nds ga BT <
10 eV . For the allowed range of axion masses
this mixing angle is too small to yield a signicant conversion eect.
Therefore, this entire line of argument is only relevant for massless
(or at least very low-mass) pseudoscalars which again shall be referred
to as arions. Carlson (1995) considered the star -Ori (Betelgeuse),
a red supergiant about 100 pc away from us. He estimated its arion
luminosity from the Primako process, and compared the expected
x-ray ux with data from the HEAO-1 satellite. As a result, a new
1
11
limit of ga <
2.510 GeV emerged which is more restrictive than
the above bound from globular-cluster stars.
Carlsons argument yields an even more restrictive limit if applied
to SN 1987A. One may estimate the arion luminosity of the SN core on
the basis of the Primako process. If arions couple to quarks or electrons, the luminosity can only be higher because existing axion limits
already indicate that arions cannot be trapped by these couplings. In
order to evaluate Eq. (5.9) an average temperature of 30 MeV and an
average density of 31014 g cm3 with a proton fraction of 0.3 is used
(Sect. 13.4.2). The Debye screening scale by the protons is then found
to be 36 MeV so that 2 = 1.41 in Eq. (5.10) leading to F = 0.72.
2
Therefore, the average energy loss rate is about g10
1.41016 erg g1 s1 .
Taking a core mass of 1M = 21033 g and a duration of 3 s one ex2
2
pects about g10
1050 erg to be emitted in arions which is about 103 g10
of the energy emitted in each neutrino avor.
Typical arion energies are 3T 100 MeV so that Eq. (5.37) together
with P 1011 eV in the interstellar medium reveals that mixing is
nearly maximal for the relevant circumstances. Therefore, the oscillation length is given by osc = 4/ga BT which is about 40 kpc for
BT = 1 G and ga = 1010 GeV1 . Therefore, osc far exceeds the
relevant magnetic eld region which is of order 1 kpc as discussed in
Sect. 13.3.3b. The conversion rate is then
prob(a ) = ( 21 ga BT )2
2
= 2.3102 g10
(BT /G kpc)2 ,
(5.38)
190
Chapter 5
than about 109 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 nds a limit of g10
1
4
11
10 or ga <
10 GeV , applicable if the particle mass is below
about 1010 eV. This is more restrictive than Carlsons original limit,
and of the same order as the PSR 1937+21 birefringence limit quoted
after Eq. (5.36).
5.5.4
Nambu-Goldstone bosons a are by denition the result of a spontaneously broken global symmetry. The cosmic evolution from a very
hot initial phase begins with the unbroken symmetry; as the universe
expands and cools a phase transition will occur where the eld responsible for the spontaneous breakdown must nd its new minimum. As
this process occurs independently in each causally connected region of
the universe at that time, the universe today will be characterized by
dierent orientations of the ground state, i.e. by dierent values of a
classical background a eld. If no ination occurred in the universe
after the phase transition, and if the Nambu-Goldstone bosons remain
truly massless (in contrast with axions), the background eld will not
have relaxed to a common ground state everywhere.
In this scenario a radio signal from a distant source travels through
regions with dierent values of the classical a eld, and thus through
regions of gradients a which act as an optically active medium according to Eq. (5.35). Therefore, linearly polarized light will experience a
random rotation of its plane of polarization.
This eect is also expected from the Faraday rotation caused by
intervening magnetic elds which induce optical activity in the cosmic
background plasma. As this eect is frequency dependent it can be
removed by observing a given object at dierent wavelengths. The
eect induced by pseudoscalars, on the other hand, is independent of
frequency.
A systematic correlation between the geometric shape of distant
radio sources and the linear polarization of the emitted radiation has
been observed. This correlation proves that no random rotation of
the plane of polarization occurs over cosmic distances, except for the
Faraday eect which can be removed from the data. Therefore, the
maximum allowed coupling strength of photons to a random cosmic
Nambu-Goldstone eld can be constrained. Harari and Sikivie (1992)
found that Ca <
50 in Eq. (5.5), independently of the symmetry break-
191
5.6
Summary of Constraints on ga
192
Chapter 5
Chapter 6
Particle Dispersion and
Decays in Media
Dispersion eects in media have a signicant impact on the propagation of some low-mass particles (photons, neutrinos) while others are
left unaected (axions and other Nambu-Goldstone bosons). The relationship between forward scattering and refraction is derived, and the
dispersion relations for photons and neutrinos are thoroughly studied.
Modied particle dispersion relations allow certain decay processes to
occur in media that cannot occur in vacuum, notably the photon decay which dominates the neutrino emissivity in a wide range
of temperatures and densities (plasma process). Other examples are
the neutrino and majoron decay and , respectively.
The rates for such processes are derived. The plasma process allows
one to derive the most restrictive limits on neutrino magnetic dipole
moments. Screening eects in reactions involving Coulomb scattering,
and neutrino electromagnetic form factors in media are discussed.
6.1
Introduction
194
Chapter 6
frequencies; m has the usual interpretation of a particle mass. One consequence of this covariant dispersion relation is that decays of the sort
1 2 + 3 are only possible if m1 > m2 + m3 so that in the rest frame of
particle 1 there is enough energy available to produce the nal states.
In media the dispersion relations are generally modied by the coherent interactions with the background. In the simplest case a particle acquires a medium-induced eective mass. For example, photons in
a nonrelativistic plasma acquire a dispersion relation 2 = P2 +k2 with
the plasma frequency given by P2 = 4 ne /me (electron density ne ).
For P > 2m this implies that the decay becomes kinematically possible and occurs in stars because the ambient electrons mediate
an eective neutrino-photon interaction (Adams, Ruderman, and Woo
1963). In fact, this plasma process is the dominant neutrino source in
a wide range of temperatures and densities which covers, for example,
white dwarfs and red-giant stars (Appendices C and D).
Neutrinos may have nonstandard electromagnetic couplings, notably magnetic dipole moments, which would enhance the plasma process and thus the cooling of stars (Bernstein, Ruderman, and Feinberg
1963). Observational constraints on anomalous cooling rates derived
from white dwarfs and globular-cluster stars then provide the most
restrictive limits on neutrino electromagnetic couplings (Sect. 6.5.6).
Within the standard model all fermions are fundamentally massless;
they acquire an eective mass by their interaction with the vacuum
expectation value 0 of a scalar Higgs eld (Sect. 8.1.1). Therefore,
even vacuum masses can be interpreted as refractive phenomena.
Because the scalar 0 is Lorentz invariant the dispersion relation thus
induced is of the standard form E 2 = m2 + p2 . Normal media,
however, single out a preferred Lorentz frame, usually causing E(p) to
be a more complicated function than (m2 + p2 )1/2 .
Notably, the dispersion relation can be such that the four-momentum P = (E, p) is space-like, P 2 = E 2 p2 < 0, which amounts
to a negative mass-square P 2 = m2e < 0. There is nothing wrong
with such tachyons because the speed of signal propagation safely
remains below the speed of light (Sect. 6.2.2). The dispersion relation
in isotropic media is often expressed as k = |k| = n in terms of a
refractive index n. Space-like excitations correspond to n > 1; examples
are photons in water or air. In this case the well-known decay process
e e is kinematically allowed for suciently fast moving electrons
(Cherenkov radiation).
The dispersion relation can also depend on the spin polarization of
the radiation. In optically active media, the left- and right-handed
195
196
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 dispersion or collective eects a kinetic approach seems often physically
more transparent while yielding identical results. At any rate, the following discussion is based entirely on kinetic theory.
6.2
6.2.1
How does one go about to calculate the all-important dispersion relation for a given particle in a medium with known properties? Usually
it is enough to follow the elementary approach of calculating the forward scattering amplitude of the relevant eld excitations with the constituents of the background medium, an approach which has the added
advantage of physical transparency over a more formal procedure.30
To begin, consider a scalar eld which may be viewed as representing one of the photon or electron polarization states. If a plane
wave excitation of that eld with a frequency and a wave vector k interacts with a scatterer at location r = 0 an additional spherical wave
will be created. The asymptotic form of the original plus scattered
wave is
(
it
(r, t) e
ikr
eikr
+ f (, )
r
(6.1)
where k = |k|, r = |r|, and f is the scattering amplitude. It was assumed that it has no azimuthal dependence, something that will always
apply on average for a collection of randomly oriented scatterers. The
dierential scattering cross section is d/d = |f (, )|2 .
If there is a collection of scattering centers randomly distributed
in space, all of the individual scattered waves will interfere. However,
because of the random location of the scatterers, constructive and destructive interference terms will average to zero. Thus the total cross
section of the ensemble is the (incoherent) sum of the individual ones.
In the forward direction, however, the scattered waves add up coherently with each other and with the parent wave, leading to a phase
shift and thus to refraction. This is seen if one considers a plane wave in
the z-direction incident on an innitesimally thin slab (thickness a) at
30
197
iz
+ na
2 1/2
eik( +z )
f (, ) 2 d ,
(2 + z 2 )1/2
(6.2)
(z) e
iz
2na
1+i
f0 () ,
(6.3)
where a term of order (z)1 was neglected which becomes small for
large z. Here, f0 () f (, 0) is the forward scattering amplitude.
Turn next to a slab of nite thickness a. The phase change of the
transmitted wave is obtained by compounding innitesimal ones with
a = a/j and taking the limit j ,
]j
2na
lim 1 + i
f0 ()
j
j
= ei(2n/)f0 a .
(6.4)
Inserting this result in Eq. (6.3) reveals that over a distance a in the
medium the wave accumulates a phase einrefr a where
nrefr = 1 +
2
n f0 ()
2
(6.5)
4
n f0 (k),
k2
(6.6)
198
Chapter 6
only refers to scattering in the forward direction, but that all properties
of the wave and the scatterer are left unchanged. If the medium particles have a distribution of momenta, spins, etc. the forward scattering
amplitude must be averaged over those quantities, and dierent species
of medium particles must be summed over.
For a practical calculation it helps to recall that d/d = |f ()|2
so that |f0 | is the square root of the forward dierential cross section.
For example, the Thomson cross section for photons interacting with
nonrelativistic electrons is d/d = (/me )2 | |2 with the polarization vectors and of the initial- and nal-state photon. Forward
scattering implies | |2 = 1 so that |f0 | = /me . The dispersion relation is then 2 = k 2 + P2 with the plasma frequency P2 = 4 ne /me .
Of course, the absolute sign of f0 has to be derived from some other
informationfor photon dispersion see Sect. 6.3.
The forward scattering amplitude and the refractive index are generally complex numbers. Physically it is evident that in a medium the
intensity of a beam is depleted as ez/ . The mean free path is given by
1 = nv where 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 eikzz/2 . Moreover, the derivation
of the refractive index indicates that the amplitude varies according to
einrefr z , yielding k = Re nrefr and (2)1 = Im nrefr . For relativistic
propagation (v = 1) the last equation implies () = (4/) Im f0 (),
a relationship known as the optical theorem.
For the applications discussed in this book specic interaction models 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 has a number of general properties, independently of the
interaction model. For example, its real and imaginary part are connected by the Kramers-Kronig relations (Sakurai 1967; Jackson 1975).
6.2.2
199
tempted to interpret p = h
k as the particles momentum. In vacuum a
particles velocity is p/E, a quantity which exceeds the speed of light
for space-like excitations. Occasionally one reads in the literature that
for this reason only those branches of a particle dispersion relation were
physical where |p| < E. Such statements are incorrect, however, and
the underlying concern about tachyonic propagation is unfounded.
The quantity p/E has no general physical relevance. Two signicant velocity denitions are the phase velocity and the group velocity of
a wave (Jackson 1975). The former is the speed with which the crest of
a plane wave propagates, i.e. it is given by the condition t kz = 0 or
vphase = /k = n1
refr . For a massive particle in vacuum vphase > 1. However, the phase velocity can drop below the speed of light in a medium.
When this occurs for electromagnetic excitations in a plasma, electrons
can surf in the wave which thus transfers energy at a rate proportional to the ne structure constant (Landau 1946), an eect known
as Landau damping. As long as vphase > 1 the photon propagation is
damped only by Thomson scattering which is an eect of order 2 .
The group velocity vgroup = d/dk is the speed with which a wave
packet or pulse propagates. In terms of the refractive index it is
1
vgroup
= nrefr () + dnrefr /d
(6.7)
200
Chapter 6
6.2.3
Wave-Function Renormalization
201
In order to determine this factor from the dispersion relation consider a scalar eld in the presence of a medium which induces a refractive index. This means that the Klein-Gordon equation in Fourier
space, including a source term , is of the form
[K 2 + (K)](K) = g(K),
(6.8)
(6.9)
(6.10)
2 + k2 + k (k )( k ) k () = gk (),
(6.11)
where the dispersion relation Eq. (6.9) was used. To rst order in k
one may use 2 = 2k = + k which allows one to write
Z 1 ( 2 k2 ) k () = gk (),
(6.12)
where
(, k)
2k k (k )
=1
.
2k
2 2 k2 =(,k)
(6.13)
202
Chapter 6
(6.14)
6.3
Photon Dispersion
6.3.1
Maxwells Equations
For the astrophysical applications relevant to this book the photon refractive index in a fully ionized plasma consisting of nuclei and electrons
will be needed. On the quantum level, this system is entirely described
by quantum electrodynamics (QED). It is sometimes referred to as a
QED plasmain contrast with a quark-gluon plasma which is described
by quantum chromodynamics (QCD). The calculation of the refractive
index amounts to an evaluation of the forward scattering amplitude of
photons on electrons, a simple task except for the complications from
the statistical averaging over the electrons which are partially or fully
relativistic and exhibit any degree of degeneracy. Recently Braaten and
Segel (1993) have found an astonishing simplication of this daunting
problem (Sect. 6.3.4).
A more conceptual complication is the occurrence of a third photon
degree of freedom in a medium (Langmuir 1926), sometimes referred to
203
= J,
BE
B = 0,
= 0.
E+B
(6.15)
(6.16)
F = 0,
(6.17)
where F is the antisymmetric eld-strength tensor with the nonvanishing components F 0i = F i0 = Ei , and F ij = F ji = ijk Bk .
Applying to the inhomogeneous equation and observing that F
is antisymmetric and symmetric under reveals that for
consistency J must obey the continuity equation.
An equivalent formulation arises from expressing the eld strengths
in terms of a four-potential A = (, A) by virtue of
F = A A ,
(6.18)
(6.19)
= = = t2 2 .
204
Chapter 6
The eld-strength tensor contains six independent degrees of freedom, the E and B elds. The redundancy imposed by the constraint of
the homogeneous equations was removed by introducing the vector potential. There still remains one redundant degree of freedom related to
a constraint imposed by current conservation. The Maxwell equations
remain invariant under a gauge transformation A A where
is an arbitrary scalar function. The modied A yields the same elds
E and B which are the physically measurable quantities.
The relationship to current conservation is easiest recognized if one
recalls that Maxwells equations can be derived from a Lagrangian
14 F 2 J A where F 2 = F F . A gauge transformation introduces
an additional term J which is identical to a total divergence (J)
if J = 0 and thus leaves the Euler-Langrange equations unchanged.
Indeed, current conservation is a necessary and sucient condition for
the gauge invariance of the theory (Itzykson and Zuber 1983).
A judicious choice of gauge can simplify the equations enormously.
Two important possibilities are the Lorentz gauge and the Coulomb,
transverse, or radiation gauge, based on the conditions
A = 0,
Lorentz gauge,
A = 0,
Coulomb gauge.
(6.20)
A = J,
Lorentz gauge,
A = JT ,
Coulomb gauge,
(6.21)
205
particles which constitute the currents move themselves under the inuence of electromagnetic elds. Therefore, the interaction between
elds and currents must be calculated self-consistently. If the elds are
suciently weak one may assume that the reaction of the currents to
the elds 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 relationship between elds and currents is nonlocal; for example, a solution
of Maxwells 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
Jind
= A .
(6.22)
(K 2 g + K K + )A = Jext
.
(6.23)
(6.24)
(6.25)
206
Chapter 6
eg K/ K 2 .
(6.26)
Next, one chooses a vector which is longitudinal relative to the spatial
207
(k 2 , k)
K K 2 U
=
,
k K2
k K2
(6.27)
where the second expression refers to the medium rest frame. There
remain two directions orthogonal to eg and eL , or equivalently, to K
and U . If k is taken to point in the z-direction two possible choices are
the unit vectors ex and ey , respectively. However, in order to retain the
azimuthal symmetry around
the k direction the circular polarization
vectors e = (ex iey )/ 2 are needed. Then
e (0, e )
(6.28)
(a = , L).
(6.29)
a Pa ,
(6.30)
a=,L
208
Chapter 6
The homogeneous Maxwell equations in Lorentz gauge in an isotropic medium then have the most general form
(
K 2 g +
a Pa A = 0.
(6.31)
a=,L
(6.32)
It yields the frequency k for modes with a given polarization and wave
number. The so-called eective mass is then m2e = a (k , k). This
expression is dierent for dierent polarizations and wave numbers,
and may even be negative.
Generally, an isotropic medium is characterized by three dierent
response functions because the left- and right-handed circular polarization states may experience dierent 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
reection because the sugar molecules have a denite 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 denes T + = and PT = P+ + P which projects on the
plane transverse to K and U in Minkowski space.
In macroscopic electrodynamics the medium eects are frequently
stated in the form of response functions to applied electric and magnetic elds instead of a response to A. The displacement induced by an
applied electric eld is D = E with the dielectric permittivity. Similarly, the magnetic eld is H = 1 B for an applied magnetic induction
where is the magnetic permeability. For time-varying and/or inhomogeneous elds these relationships are understood in Fourier space
where the response functions depend on and k.
The magnetic eld 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 1 ) k 2 / 2 the transverse dielectric permittivity.
209
and
T = 1 T / 2 .
(6.33)
and
2 T (, k) = k 2
(6.34)
L = e
L eL and T = e e , or explicitly in the medium frame
(Weldon 1982a)
L = (1 2 /k 2 ) 00
and
T = 12 (Tr L ),
(6.35)
with Tr = g .
Recall that the dispersion relations are given by 2 k 2 = T,L (, k).
Thus the frequency (k) and the eective mass of a given mode are
generally complicated functions of k, notably in a medium involving
bound electrons where various resonances occur. It can be shown on
general grounds (Jackson 1975), however, that for frequencies far above
all resonances the transverse mode has a particle-like dispersion relation
2 k 2 = m2T where mT is the transverse photon mass which is a
constant independent of the wave number or frequency.
6.3.4
210
Chapter 6
formalism is not required because the only contribution is from lowestorder forward scattering on charged particles. Moreover, because the
scattering amplitude involves nonrelativistically the inverse mass of the
targets one may limit ones attention to the electrons.
Then one takes the standard (truncated) Compton-scattering matrix element (e.g. Bjorken and Drell 1964; Itzykson and Zuber 1983)
and takes an average over the Fermi-Dirac distributions of the electrons. To lowest order in = e2 /4 this yields (Altherr and Kraemmer
1992; Braaten and Segel 1993)
(K) = 16
d3 p
2E(2)3
1
e(E)/T + 1
e(E+)/T + 1
(P K)2 g + K 2 P P P K (K P + K P )
,
(P K)2 14 (K 2 )2
(6.36)
where P = (E, p) and E = (p2 + m2e )1/2 , apart from refractive eects
for the electrons and positrons. The phase-space distributions represent
electrons and positrons at temperature T and chemical potential .
Over the years, the phase-space integration has been performed in
various limits (Silin 1960; Tsytovich 1961; Jancovici 1962; Klimov 1982;
Weldon 1982a; Altherr, Petitgirard, and del Ro Gaztelurrutia 1993).
The most comprehensive analytic result is that of Braaten and Segel
(1993) which contains all previous cases in the appropriate limits.
The main simplication occurs from neglecting the (K 2 )2 term in
the denominator of Eq. (6.36). For light-like Ks this is exactly correct, and in the nonrelativistic limit where me is much larger than all
other energy scales the approximation is also trivially justied. In the
relativistic limit it is only justied if one is interested in (K) near the
light cone ( = k) in Fourier space. In the relativistic limit both transverse and longitudinal excitations have dispersion relations which are
approximately ( 2 k 2 )1/2 e T or eEF in the nondegenerate and degenerate limits, respectively. As detailed by Braaten and Segel (1993),
this deviation from masslessness is small enough to justify the approximation if one aims at the dispersion relations. Including 14 (K 2 )2 yields
an O(2 ) correctionit can be ignored in an O() result.
211
In a higher-order calculation one has to include the proper e dispersion relations which imply that electromagnetic excitations are never
damped by e+ e decay (Braaten 1991), in contrast with statements found in the previous literature. Dropping the (K 2 )2 term in the
denominator of Eq. (6.36) prevents from developing an imaginary
part from this decay, even with the vacuum e dispersion relations.
Therefore, the approximate integral actually provides a better representation of the O() dispersion relations than the exact one.
With the approximation K 2 = 0 in the denominator of Eq. (6.36)
the angular integral is trivial.32 With Eq. (6.35) one nds
[
(
)
]
p2
2 k2
4 2 k 2
+ kv
dp fp
2
1 ,
L =
log
k2
E kv
kv
k2v2
0
[
(
)]
p2
4 2 k 2
2
+ kv
dp fp
T =
log
, (6.37)
k2
E 2 k 2 2kv
kv
0
L = P2 1 G(v2 k 2 / 2 ) + v2 k 2 k 2 ,
[
T = P2 1 + 12 G(v2 k 2 / 2 ) .
(6.38)
G(x) =
1
log
x
3
2 x
1 x
=6
xn
.
n=1 (2n + 1)(2n + 3)
(6.39)
(6.40)
(6.41)
In the degenerate limit, Jancovici (1962) has calculated analytically the full
integral without the K 2 = 0 approximation.
212
Chapter 6
Dispersion Relations
In order to determine the photon dispersion relation for specic conditions one must determine P and v corresponding to the temperature
T and chemical potential of the electrons. In Fig. 6.3 contours for
v and P /T are shown in the T --plane of a plasma. Analytic
limiting cases are (Braaten and Segel 1993)
1/2
(5T /me )
v =
vF
(
)
4 ne
5 T
me
2 me
4 n
4 2
e
P2 =
=
p F vF
E
3
F
4 ( 2 1 2 2 )
+ T
Classical,
Degenerate,
Relativistic,
(6.42)
Classical,
Degenerate,
(6.43)
Relativistic,
3
3
where vF = pF /EF is the velocity at the Fermi surface, classical refers
to the nondegenerate and nonrelativistic limit, and relativistic is for
any degree of degeneracy.
213
Fig. 6.3. Contours for v and = P /T as dened 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
T,L (, k) = 2 k 2 which are explicitly
[
2 k2
= P2 1 + 12 G(v2 k 2 / 2 )
[
2 v2 k 2 = P2 1 G(v2 k 2 / 2 )
Transverse,
Longitudinal.
(6.44)
P2
P2
=k +
(
k2 T
1+ 2
me
k2 T
1+3 2
me
Transverse,
Longitudinal.
(6.45)
For small temperatures the longitudinal modes oscillate with an almost xed frequency, independently of momentum, while the transverse
modes behave almost like massive particles (Fig. 6.4).
The general result Eq. (6.38) and the behavior of the function G(x)
reveal that for transverse excitations 2 k 2 can vary only between P2
and 23 P2 . Also, 2 k 2 > 0 so that K 2 is always time-like. For k P
the transverse dispersion relation approaches that of a massive particle
with a xed mass mT , the transverse photon mass. With Eq. (6.38)
and because k/ 1 for k P one nds m2T = P2 [1 + 12 G(v2 )], or
214
Chapter 6
4
p2
=
dp fp .
0
E
(6.46)
1 + vF
1 vF2
m2T
1
log
=
2
2
2v
2vF
1 vF
P
3
2
)]
Classical,
Degenerate,
(6.47)
Relativistic.
215
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).
p2 1
1+v
dp fp
log
1
E v
1v
0
[
(
)
]
1
1 + v
2 3
= P 2
log
1 .
v 2v
1 v
4
=
(6.48)
The second identity (Braaten and Segel 1993) applies at the same level
of approximation as T,L in Eq. (6.38). Some analytic limiting cases
216
Chapter 6
are
k12
P2
1 + 3T /me
1
1 + vF
log
1
2vF
1 vF
vF2
Classical,
(6.49)
Degenerate,
Relativistic.
kD
k
)3
D
e 2k2 =
k2
( )3/2 (
5
2
P
v k
)3
5 P
e 2 v2 k2 , (6.50)
2
=
where kD = 4 ne /T is the Debye screening scale. (Note that P2 /kD
T /me = v2 /5.) For a given wave number a plasmon must be viewed
217
Armed with the dispersion relation one may determine the vertex renormalization constants ZT,L relevant for the coupling of external photons
or plasmons to an electron in the medium (Sect. 6.2.3),
1
ZT,L
T,L (, k)
=1
.
2 2
2
k =T,L (,k)
(6.51)
2 2 ( 2 v2 k 2 )
,
2 [3P2 2 ( 2 k 2 )] + ( 2 + k 2 )( 2 v2 k 2 )
ZL =
2
2 ( 2 v2 k 2 )
.
3P2 ( 2 v2 k 2 ) 2 k 2
(6.52)
In each case and k are on shell, i.e. they are related by the dispersion 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 massiveparticle dispersion relation. The contours in Fig. 6.7 conrm 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. Modied vertex renormalization factor ZeL for longitudinal electromagnetic excitations in a medium according to Eq. (6.53).
2 ( 2 v2 k 2 )
,
3P2 ( 2 v2 k 2 )
(6.53)
219
6.4
Screening Eects
6.4.1
Debye Screening
(6.54)
220
Chapter 6
(t J = 0) are not screened. The magnetic eld associated with a stationary current is the same at a distance whether or not the plasma
is present.
Not so for the electric eld associated with a charge. In the static
limit one nds
4
L (0, k) =
dp fp p (v + v 1 ).
(6.55)
(6.56)
(6.57)
4 ne
me 2
=
.
T
T P
(6.58)
2
At this point one recognizes that kD
is independent of the electron
mass, in contrast with the plasma frequency P2 . Therefore, it is no
longer justied to ignore the ions or nuclei; they contribute little to
dispersion because of their reduced Thomson scattering amplitude, but
2
they contribute equally to screening. Therefore, one nds kS2 = kD
+ ki2
with
ki2 =
4
nj Zj2 ,
T
j
(6.59)
221
4
3 2
EF pF = 2P .
vF
(6.60)
2
2
so that in a medium of degenerate
always exceeds kTF
However, kD
electrons and nondegenerate ions the main screening eect is from the
latter. Recall that the Fermi momentum is related to the electron
density by ne = p3F /3 2 and the Fermi energy is EF = (p2F + m2e )1/2 .
To compare the Thomas-Fermi with the Debye scale take the nonrelativistic limit (kTF /kD )2 = 32 T /(EF me ). This is much less than 1
or the medium would not be degenerate whence kTF kD . Therefore, if the electrons are degenerate and the ions nondegenerate, a test
charge is mostly screened by the polarization of the ion uid because
the electrons form a sti background. Unfortunately, one often nds
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 nal answer (see below).
Screening eects in Coulomb processes are often found to be implemented by a modied Coulomb propagator
1
1
,
|q|4
(q2 + kS2 )2
(6.61)
For a textbook derivation from a Thomas-Fermi model see Shapiro and Teukolsky (1983). Note that they work in the nonrelativistic limit: their EF = p2F /2me .
222
Chapter 6
The screening of electric elds in a plasma is closely related to correlations of the positions and motions of the charged particles. If a negative
test charge is known to be in a certain position, the probability of nding an electron in the immediate neighborhood is less than average,
while the probability of nding a nucleus is larger than average. It is
this polarization of the surrounding plasma which screens a charge.
Take one particle of a given species to be the origin of a coordinate
system, and take their average number density to be n. The electrostatic repulsion of the test charge causes a deviation of the surrounding
charges from the average density by an amount
S(r) = 3 (r) + n h(r),
(6.62)
S(q) =
d3 r S(r) eiqr
(6.63)
kS2 ekS r
.
4 r
(6.64)
The volume integral of (r) vanishes, giving zero total charge, i.e. complete screening at innity. 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 /4r) ekS r . As expected, Debye screening corresponds
223
to spatial anticorrelations of like-charged particles. Fourier transforming Eq. (6.64) yields the important result
S(q) =
q2
q2 + kS2
(6.65)
2
q2 + ZkD
,
2
q2 + (1 + Z)kD
Sii (q) =
2
q2 + kD
,
2
q2 + (1 + Z)kD
Sei (q) =
2
kD
.
2
q2 + (1 + Z)kD
(6.66)
q2
,
q2 + kS2
(6.67)
2
2
with kS2 = kD
+ ki2 = (1 + Z) kD
. (Note that for only one species of ions
2
2
ki = ZkD .) Hence one reproduces a screened charge distribution which
causes a Yukawa potential. However, the small-q behavior of See or Sii
is very dierent: See (0) = Z/(1 + Z) in a two-component plasma while
See (0) = 0 for only one component.
6.4.3
For low temperatures, the screening will not be of Yukawa type and
the structure factor will deviate from the simple Debye formula. A
plasma can be considered cold if the average Coulomb interaction energy between ions is much larger than typical thermal energies. To
quantify this measure, one introduces the ion-sphere radius ai by
= 4a3i /3 where ni is the number density of the mobile
virtue of n1
i
224
Chapter 6
=
(6.68)
ai T
3
as a measure for how strongly the plasma is coupled, where ki2 =
4Z 2 /T . For 1 it is weakly coupled and approaches an ideal
Boltzmann gas.
The Debye structure factor of a one-component plasma can be written as
|ai q|2
SD (q) =
.
(6.69)
|ai q|2 + 3
This result applies even for large if |ai q| 1. For 1, the plasma
is strongly coupled, and for >
178 the ions will arrange themselves
in a body centered cubic lattice (Slattery, Doolen, and DeWitt 1980,
1982).
In Fig. 6.9 I show S and SD as functions of ai q = |ai q| for = 2, 10
and 100 where S was numerically determined (Hansen 1973; Galam and
Hansen 1976). The emerging periodicity for a strongly coupled plasma
is quite apparent. It is also clear that for <
1 the Debye formula gives
a fair representation of the structure factor while for a strongly coupled
plasma it is completely misleading. The interior of white dwarfs is in
the regime of large , and old white dwarfs are believed to crystallize.
(See Appendix D for an overview over the conditions relevant for stellar
plasmas.)
6.4.4
Armed with these insights one may turn to the issue of Coulomb scattering processes in a plasma. In the limit of nonrelativistic and essentially
static sources for the electric elds the relevant quantity entering the
matrix element is the Fourier component (q) of the charge distribution (r) where q is the momentum transferred by the Coulomb eld
to the sources. The squared matrix element thus involves the quantity
(q) (q) which is (q)(q) because (r) is real. Taking a statistical
average over all possible congurations of the charge distribution leads
to a rate proportional to
S(q) = (q)(q).
(6.70)
225
226
Chapter 6
(6.71)
which implies
1
1
2 2
4
|q|
q (q + kS2 )
(6.72)
+1
1
dx
(1 x) f (x)
,
(1 x)2
(6.73)
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
+1
1
2 + 2
1
= log
,
dx
(1 x + 2 )
2
(1 x)
2 + 2
dx
=
log
(1 x + 2 )2
2
2
,
2 + 2
(6.74)
where 2 kS2 /2p2 is the screening scale expressed in units of the initialstate momentum of the probe. Usually, it far exceeds the screening scale
whence 2 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).
6.5
6.5.1
227
(6.75)
|M|2 = M P P
(6.76)
spins
where explicitly
M = 4e2 Z (g + 2 ).
(6.77)
228
Chapter 6
d3 p
d3 p
1
4 4
(2)
(K
P
)
|M|2 .
3
3
2Ep (2) 2Ep (2)
2 spins
(6.78)
)
( 2
d3 p d3 p 4
P P (K P P ) =
K g + 2K K .
2Ep 2Ep
24
(6.79)
(6.80)
where the normalization = 1 for transverse and time-like longitudinal plasmons was used as well as K = 0. applies to both
transverse and longitudinal plasmons with the appropriate ZT,L . For a
chosen three-momentum k = |k| the quantities Z, , and K 2 = 2 k 2
are all functions of k by virtue of the dispersion relation K 2 = T,L (K).
In the classical limit transverse plasmons propagate like massive
particles with K 2 = P2 and ZT = 1. Then T = 31 P3 (P /) where
the last factor is recognized as a Lorentz time-dilation factor. For a
general dispersion relation which is not Lorentz covariant it makes little
sense, of course, to express the decay rate in the plasmon frame. An
example is the classical limit for the longitudinal mode for which to
zeroth order in T /me the frequency is = P , ZL = 2 /K 2 , and then
L = 13 P with the restriction k < P .
6.5.2
Another direct coupling between neutrinos and photons arises if the former have electric or magnetic dipole or transition moments (Sects. 7.2.2
and 7.3.2)
Lint =
1
2
ab a b + ab a 5 b F .
(6.81)
a,b
229
The squared matrix element is of the form Eq. (6.76); for a magnetic
dipole coupling of a single avor one nds
M = 42 Z (2K K 2K 2 K 2 g ).
(6.82)
(6.83)
|ab |2 + |ab |2 .
(6.84)
a,b
There is no interference term between the electric and magnetic couplings. The presence of transition moments allows for plasmon decays
of the sort a b with dierent avors a = b, doubling the nal
states which may be a b or b a .
6.5.3
Standard-Model Couplings
Fig. 6.11. Compton production of neutrino pairs (nonzero momentum transfer to electrons) or plasmon decay (electron forward scattering).
230
Chapter 6
With the dimensionless coupling constants CV and CA given in Appendix B the neutral-current -e-interaction is
GF
Lint = e (CV CA 5 )e (1 5 ) .
2
(6.85)
The vector-current has the same structure that pertains to the electron interaction with photons, Lint = ie e e A . Therefore, after 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 forward scattering is the same as the conversion .
231
present discussion I will not worry any further about the axial-vector
contribution.
The matrix element for the interaction between neutrinos and photons can then be read from the eective vertex
CV GF
i A (1 5 ) .
e 2
(6.86)
The electric charge in the denominator removes one such factor contained in which was calculated for photon forward scattering. In
the matrix element, is to be taken at the four-momentum K of the
photon. Besides plasmon decay , this interaction also allows for
processes such as Cherenkov absorption or emission .
For the decay of a plasmon with a polarization vector and fourmomentum K the squared matrix element has the form Eq. (6.76) with
M = 8
G2F 2
(g + 2 ).
2e2 T,L
(6.87)
CV2 G2F
( 2 k 2 )3
Z
.
T,L
48 2
(6.88)
This result was rst derived by Adams, Ruderman, and Woo (1963),
the correct Z for the longitudinal case was rst derived by Zaidi (1965).
This equation is understood on shell where depends on k through
the dispersion relation 2 k 2 = T,L (, k).
6.5.4
T,L (k)
T,L (k , k)
,
P2
(6.89)
232
Chapter 6
The decay rates of Eqs. (6.80), (6.83) and (6.88) of a plasmon with
three-momentum k are then expressed as
4 Zk
k =
3 k
2
2
P2
k
4
P2
k
4
Millicharge,
)2
Dipole Moment,
(
)3
CV2 G2F P2
k
(6.90)
Standard Model,
Energy-Loss Rates
(6.91)
83 3
Q=
T
2
2
P2
4
P2
4
Q1
Millicharge,
Q2
Dipole Moment,
)2
(
)3
CV2 G2F P2
Q3
(6.92)
Standard Model,
where 3 1.202 refers to the Riemann Zeta function. The dimensionless emission rates Qn for the three cases are each a sum of a transverse
233
2 e
1 k1
1 k1
Ln1
Ln
2 ZL
2 ZL
=
=
,
dk
k
dk
k
43 T 3 0
e/T 1
43 T 3 0
P2 e/T 1
Tn
1
2 ZT
dk
k
=
.
23 T 3 0
e/T 1
(6.93)
The second equation for QL,n relies on the denition Eq. (6.53), i.e.
ZL = ZeL 2 /( 2 k 2 ), and 2 k 2 = L was used.
The normalization factors were chosen such that QT,n = 1 if the
plasmons are treated as eectively massless particles for the phase-space
integration. Then ZT =
T = 1 which is a reasonable approximation
in a nondegenerate, nonrelativistic plasma. In that limit to lowest
order k1 = P , ZeL = 1, and L = P2 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, Raelt, 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 between Fig. C.1 and Fig. 6.3 reveals that the plasma process is important
<
for 0.3 <
P /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 neutrino 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 nds
(4)2 Q1
Qcharge
= 2 2 4
= 0.664 e214
QSM
CV GF P Q3
Qdipole
2 2 Q2
= 2 2 2
= 0.318 212
QSM
CV GF P Q3
10 keV
P
10 keV
P
)4
)2
Q1
,
Q3
Q2
,
Q3
(6.94)
234
Chapter 6
Fig. 6.13. Contours of Q1 /Q3 and Q2 /Q3 dened by Eq. (6.93) in the plane
dened by the plasma frequency P and a typical electron velocity v
discussed in Sect. 6.3. See Fig. 6.3 for contours of v and P in the T -plane. (Adapted from Haft, Raelt, and Weiss 1994.)
6.5.6
235
236
Chapter 6
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 rst of helium, later
also of carbon and oxygen, for all of which Ye = 0.5. Then,
5.0 e214
0.0127 24
Millicharge,
Dipole Moment, (6.95)
Standard Model,
and
12
<
1410 B .
(6.96)
Of course, for such large dipole moments the core would grow far beyond 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 eect as shown, for example, by the
numerical calculations of Raelt, 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 heliumignition argument where 2105 g cm3 . According to Eq. (D.12)
the relevant plasma frequency is P = 8.6 keV so that Qcharge /QSM
1.2 e214 and Qdipole /QSM 0.4 212 in Eq. (6.94). The average total emission rate is then given by the standard rate times F = 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
14
than 5% one must require F < 3. Then one nds e <
1.310 e
12
and <
210 B . For the dipole case, a detailed numerical imple12
mentation yielded <
310 B (Sect. 2.5.2), nearly identical with
this simple analytic estimate. The limit on the charge could also be
slightly degraded and so I adopt
14
e <
210 e
and
12
<
310 B
(6.97)
6.6
237
Le = 2 GF 12 (1 5 ) A ,
(6.98)
238
Chapter 6
V = 4e CV
]
d3 p [
+
f
(p)
+
f
(p)
e
e
2E(2)3
(P K)2 g + K 2 P P P K (K P + K P )
,
(P K)2 41 (K 2 )2
= 2ie CA
]
d3 p [
(p) fe+ (p)
f
e
2E(2)3
K 2 P K
,
(P K)2 41 (K 2 )2
(6.99)
V = (CV /e) .
(6.100)
In an isotropic plasma, is characterized by the two medium characteristics T (, k) and L (, k) which are functions of the photon fourmomentum K = (, k) with k = |k|. For the antisymmetric piece, in
an isotropic medium the phase-space integration averages the spatial
part of P to zero so that
0
K a(, k)
A = 2ieCA
(6.101)
239
this process is kinematically forbidden. However, longitudinal electromagnetic excitations (plasmons), which exist only in the medium,
propagate such that for some momenta 2 k2 < 0 (space-like fourmomentum), allowing for Cherenkov emission. One nds statements
in the literature that the neutrino energy transfer to the medium by
this process exceeded the transfer by (incoherent) -e scattering (e.g.
Oraevski and Semikoz 1984; Oraevski, Semikoz, and Smorodinski
1986; Semikoz 1987a). This is in conict with the discussion of Kirzhnits, Losyakov, and Chechin (1990) who found on general grounds that
the energy loss of a neutrino propagating in a stable medium was always bounded from above by the collisional energy loss, apart from a
factor of order unity. Granting this, the Cherenkov process does not
seem to be of great practical importance.
If neutrinos have masses and mix, decays of the form 2 1 are
possible in vacuum, and can be kinematically possible in a medium if
the photon eective mass does not exceed the neutrino mass dierence m2 m1 . If kinematically allowed, this decay receives a contribution from the medium-induced coupling which may far exceed the
vacuum decay rate. Explicit calculations were performed by a number of authors36 who unfortunately ignored the kinematic constraint
imposed by the photon dispersion relation. This is not a reasonable
approximation in view of the relatively small neutrino masses that remain of practical interest. Further, in order to judge the importance
of the medium-induced decay it is not relevant to compare with the
vacuum decay rate, but rather one should compare with the collisional
transition rate 2 e e1 (mediated by photon exchange) which is the
process with which the coherent reaction directly competes.
The photon decay as well as the Cherenkov process and the mediuminduced neutrino decay all have in common that the neutrino couples
to an electromagnetic eld which is a freely propagating wave, obeying
the dispersion relation in the medium which is 2 k 2 = T,L (, k)
for transverse and longitudinal excitations, respectively. However, one
may also consider the eect of a static external electric or magnetic
eld.37 To this end, one must take the static limit 0 of the
vertex functions
V,A (, k). For an external static electric eld the
only nonvanishing component of the vector potential A is A0 . Then
36
DOlivo, Nieves, and Pal (1990); Kuo and Pantaleone (1990); Giunti, Kim, and
Lam (1991).
37
This issue has been investigated in many works, e.g. Oraevski and Semikoz
(1985, 1987), Semikoz (1987a,b), Nieves and Pal (1989c, 1994), Semikoz and
Smorodinski (1988, 1989), and DOlivo, Nieves, and Pal (1989).
240
Chapter 6
00
(6.102)
e = (CV /e) 2 GF kS2 .
= eCA 2GF 4
dp fe (p) fe+ (p) .
(6.103)
0
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 eld. For a
rst discussion see Oraevski 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 discussed in the previous section.
6.7
6.7.1
Neutrino Refraction
Neutrino Refractive Index
g
Q2 g Q Q
+
+ ...
m2Z,W
m4Z,W
(6.104)
and keep only the rst term. (The second term is needed if the contribution of the rst one cancels as in a CP symmetric mediumsee
38
In the formalism of nite temperature and density (FTD) eld theory the amplitudes may be written in a more compact form so that the relevant Feynman
graphs reduce to a tadpole and a bubble graph (Notzold and Raelt 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 KovesiDomokos (1995) for a discussion of the oscillations of very high-energy cosmic neutrinos for which this approximation is not adequate.
242
Chapter 6
Fig. 6.15. Amplitudes contributing to forward scattering: (a) Neutralcurrent scattering for any or on any f or f as target. (b) e - e scattering.
(c) e -e charged-current scattering. (d) e -e+ charged-current scattering.
(e) Eective four-fermion vertex in the low-energy limit.
(6.105)
where is a neutrino eld ( = e, , ) while f represents fermions of the medium (f = e, p, n, or even neutrinos ). Here, GF =
1.166105 GeV2 is the Fermi constant. The relevant values of the
vector and axial-vector weak charges CV and CA are given in Appendix B. In the low-energy limit the charged-current reactions (c)
and (d) can also be represented as an eective neutral-current interaction of the same form with CV = CA = 1.
It is now straightforward to work out the forward scattering amplitudes. The axial-vector piece represents the spin of f and so it averages
to zero if the medium is unpolarized. Then one nds for the refractive
index of a neutrino (upper sign) or antineutrino (lower sign) with energy
nrefr 1 = CV GF
nf nf
,
2
(6.106)
where nf and nf are the number densities of fermions f and antifermions f , respectively. The eective weak coupling constants CV are
identical with the CV given in Appendix B except for neutrinos as
243
medium particles40 which are left-handed and thus polarized. Therefore, (1 5 ) = 2 and CV = 2CV .
Because we are dealing with forward scattering where recoil eects
do not occur, the contributions from free or bound nucleons are the
same. Therefore, Eq. (6.106) allows one to determine the refractive
index of any normal medium. Because electric neutrality implies an
excess density of electrons over positrons which balances against the
protons, their neutral-current contributions cancel. (A possible exception is a condensate that may exist in neutron stars.) An excess
of e over e appears to occur only in a young supernova core where
neutrinos have a large chemical potential for the rst few seconds after
collapse.
All told, the dispersion relation for unmixed neutrinos, valid even in
the nonrelativistic limit (Chang and Zia 1988), can be written in terms
of a potential energy as
( V )2 = k 2 + m2 ,
(6.107)
( 1 Yn + Ye + 2Ye ) for e ,
2
V = 2GF nB
( 1 Y + Y )
for , ,
e
2 n
(6.108)
nf nf
nB
(6.109)
2 GF nB = 0.7621013 eV
(6.110)
g cm3
with the mass density .
A remark concerning the absolute sign of V is in order. The relative
signs between the dierent CV s can be worked out easily from the
weak interaction structure of the standard model. Also, the relative
sign of the eective neutral-current amplitudes which follow from Z
and W exchange follows directly, for example, from the FTD approach
(Notzold and Raelt 1988). Thus to x the overall sign it is enough
40
244
Chapter 6
nrefr =
V
1
)2
m2
2
]1/2
V
m2
2
(6.111)
(6.112)
2 2GF nB
)1/2
(
4
= 3.9110
eV
g cm3
)1/2 (
MeV
)1/2
.
(6.113)
245
The absolute shift of the neutrino masses is rather negligible because we are dealing with highly relativistic particles. Even in this limit,
however, the dierence between the dispersion relation of dierent avors is important for oscillation eects. Hence the most noteworthy
medium eect is its avor birefringence: e and , acquire dierent
eective masses because of the charged-current contribution from
e -e
scattering. The dierence of their potentials is Ve V, = 2GF nL
with nL = YL nB the lepton-number density where the number fraction of leptons is YL = Ye + Ye . Of course, neutrinos as a background
medium contribute only in a young supernova core.
6.7.2
Higher-Order Eects
In the early universe one has nearly equal densities of particles and antiparticles with an asymmetry of about 109 , leading to a near cancellation of the refractive terms Eq. (6.106). One may think that the next
most important contribution is from - scattering, a process closely
related to the decay briey discussed in Sect. 7.2.2. If one
approximates the weak interactions by an eective four-fermion coupling the relevant amplitude is given by the graph Fig. 6.16 which on
dimensional grounds should be of order GF . However, electromagnetic gauge invariance together with the left-handedness of the weak
interaction implies that it vanishes identically (Gell-Mann 1961). For
massive neutrinos the amplitude is proportional to GF m , but even in
this case it vanishes in the forward direction (Langacker and Liu 1992).
246
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 ( = e, , )
in a photon bath,
GF
nrefr 1 =
4 m2W
4
m2W
1 + ln
3
m2
)]
E n ,
(6.114)
)
8 2 GF (
nrefr 1 =
E
n
+
E
n
3m2
Z
)
8 2 GF (
n + E+ n+ .
+
E
(6.115)
3m2W
In this case the contributions from background fermions and antifermions add with the same sign, and the global sign remains the same for
and as test particles. In practice, only an electron-positron background is of relevance in the early universe so that , for example,
only feels a second-order contribution from other s and s. These
results are of order G2F / and thus they are the dominant contribution
in a CP-symmetric plasma.
One-loop corrections to the amplitudes of Fig. 6.15 yield other
higher-order terms which are of order G2F like Eq. (6.114). They are
still interesting because the loops involve charged leptons with a mass
depending on their avor. Therefore, the universality of the eective
neutral-current interaction is broken on this level, leading to dierent
refractive indices for dierent . Between e and or the medium
is already birefringent to lowest order from e -e charged-current interactions. Between and the one-loop correction dominates. Assum41
2
2
Note that m2
Z = cos W sin W
2
2 GF / and m2
W = sin W
2 GF /.
247
ing electric neutrality it is (Botella, Lim, and Marciano 1987; see also
Semikoz 1992 and Horvat 1993)
[
V V
3G2F m2
m2W
=
n
ln
B
2 2
m2
Yn
1+
,
3
(6.116)
with a sign change for V V . The shift of the eective mass m2e
is numerically
(
2(V V ) = 2.06106 eV
)2
6.61 + Yn /3
,
3
g cm MeV
7
(6.117)
(6.118)
(6.119)
where is the angle relative to the local tangential vector, i.e. d/ds
is the local curvature of the beam, and is the transverse gradient.
The total angle of deection is
|| =
ds | nrefr |
(6.120)
248
Chapter 6
R
b
b r nrefr
dr 2
.
r b2
(6.121)
2 GF nB,c R b r [nB ( 21 Yn + Ye )]
= 2
dr
,
(6.122)
b
nB,c r2 b2
where nB,c is the baryon density at the center of the lens. This expression applies to e while for or the term Ye is absent, and for
antineutrinos the overall sign changes.
For the Sun with a central density of about 150 g cm3 the overall
coecient is 2.31018 (10 MeV/). The integral expression is dimensionless and thus of order unity. Therefore, the focal length of the
Sun as a neutrino lens is of order 1018 R (solar radius) for 10 MeV
neutrinos, or about the radius of the visible universe!
6.8
Majoron Decay
(6.123)
249
they are massless there is no operational distinction between a Majorana neutrino and the two active degrees of freedom of a Dirac neutrino.
Therefore, according to Eq. (6.107) the dispersion relation for the helicity states of a Majorana neutrino is
E = (m2 + p2 )1/2 V,
(6.124)
1
e it is V = 2GF (ne 2 nn ) with the electron and neutron densities
ne and nn , respectively.
This dispersion relation implies that the medium is optically active with regard to the neutrino helicities, just as some media are
birefringent with regard to the photon circular polarization. In the
optical case the left-right symmetry (parity) is broken by the medium
constituents which must have a denite handedness; sugar molecules
are a well-known example. In the neutrino case parity is broken by the
structure of the interaction; the medium itself is unpolarized.
Because there is an energy dierence between the Majorana helicity
states for a given momentum, decays + are kinematically
allowed. For relativistic neutrinos the squared matrix element is found
to be |M|2 = 4h2 P1 P2 with the four-momenta P1,2 of the initial and
nal neutrino state. The dierential decay rate is then
d =
4h2 d3 p2
d3 k
P1 P2 (2)4 4 (P1 P2 K)
2E1 2E2 (2)3 2(2)3
(6.125)
+1
P1 P2 p22
dx
(E1 E2 ),
E1 E2
(6.126)
(6.127)
250
Chapter 6
If one ignores the vacuum mass relative to V one has p1,2 = E1,2 V
so that to lowest order in V
d
V (E1 E2 )
.
=
dE2
E12
(6.128)
(6.129)
Chapter 7
Nonstandard Neutrinos
The phenomenological consequences of nonvanishing neutrino masses
and mixings and of electromagnetic couplings are explored. The decay
channels and electromagnetic properties of mixed neutrinos are discussed. Experimental, astrophysical, and cosmological limits on neutrino masses, decays, and electromagnetic properties are summarized.
7.1
7.1.1
Neutrino Masses
The Fermion Mass Problem
In the physics of elementary particles one currently knows of two categories of apparently fundamental elds: the spin- 12 quarks and leptons on the one-hand side, and the spin-1 gauge bosons on the other.
The former constitute matter while the latter mediate the electromagnetic, weak, and strong forces. The gauge-theory description of
the interactions among these particles is renowned for its elegance and
stunning in its success at accounting for all relevant measurements. At
the same time it has many entirely loose ends. Perhaps the most puzzling problem is that of fermion masses and the related issue of the
threefold replication of families: The electron and neutrino as well as
the up and down quarks (which make up protons and neutrons) each
come in two additional avors or families which seem to dier from
the rst one only in their masses (Fig. 7.1).
The standard model of particle physics holds that all fermions and
gauge bosons are fundamentally massless. The gauge symmetry forbids a fundamental mass for the latter while the masslessness of the
former is indicated by the handedness of the weak interaction: only
left-handed (l.h.) fermions feel this force while the right-handed (r.h.)
251
252
Chapter 7
Nonstandard Neutrinos
253
Neutrinos break the pattern of Fig. 7.1 in that they are much lighter
than the other members of a given family, a discrepancy which is most
severe for the third family where cosmologically m <
30 eV, eight and
ten orders of magnitude less than m and mt , respectively! Moreover,
neutrinos are dierent in that their r.h. chirality states are sterile because of the handedness of the weak interaction. The r.h. states interact
with the rest of the world only by gravity and by a possible Yukawa
coupling to the Higgs eld.
It is frequently assumed that neutrinos do not couple to the Higgs
eld, and that the r.h. components do not even exist, assumptions which
are part of the particle-physics standard model. In this case there are
only two neutrino states for a given family as opposed to four states for
the charged leptons. Actually, one may interpret the two components
of such a neutrino as the spin states of a Majorana fermion which is dened to be its own antiparticle. Fermions with four distinct states are
known as Dirac fermions. Naturally, a Majorana fermion cannot carry
a charge as that would allow one to distinguish it from its antiparticle.
A magnetic or electric dipole moment is equally forbidden: its orientation relative to the spin is reversed for antiparticles. For example, the
neutron cannot be a Majorana fermion among other reasons because it
carries a magnetic moment.
Because the r.h. neutrino components of a given family, if they exist,
are sterile anyway, there is no practical distinction between massless
Dirac and Majorana neutrinos except in a situation where gravitational
254
Chapter 7
(7.1)
Nonstandard Neutrinos
255
kinematical impact of a mass on certain reactions such as nuclear decays of the form (A, Z) (A, Z + 1) e e where the continuous energy
spectrum of the electrons originally revealed the emission of another
particle that carried away the remainder of the available energy. The
minimum amount of energy taken by the neutrino is the equivalent of
its mass so that the upper endpoint of the electron spectrum is a sensitive measure for me . Actually, the most sensitive probe is the shape
of the electron spectrum just below its endpoint, not the value of the
endpoint itself. The best constraints are based on the tritium decay
3
H 3 He+ e e with a maximum amount of kinetic energy for the
electron of Q = 18.6 keV. This unusually small Q-value ensures that a
large fraction of the electron counts appear near the endpoint (Boehm
and Vogel 1987; Winter 1991).
In Tab. 7.1 the results from several recent experiments are summarized which had been motivated by the Moscow claim of 17 eV <
me < 40 eV (Boris et al. 1987). This range is clearly incompatible
with the more recent data which, however, nd negative mass-squares.
This means that the endpoint spectra tend to be slightly deformed
in the opposite direction from what a neutrino mass would do. This
eect is particularly striking and signicant for the Livermore experiment where it is nearly impossible to blame it on a statistical uctuation. Therefore, at the present time one cannot escape the conclusion that this experimental technique suers from some unrecognized systematic eect. This problem must be resolved before it will
become possible to extract a reliable bound on me although it appears unlikely that an me in excess of 10 eV could be hidden by what-
Experiment
Los Alamos
Tokyo
Z
urich
Mainz
Livermore
Troitsk
147 68 41
65 85 65
24 48 61
39 34 15
130 20 15
18 6
a See
Reference
Robertson et al. (1991)
Kawakami et al. (1991)
Holzschuh et al. (1992)
Weinheimer et al. (1993)
Stoe and Decman (1994)
Belesev et al. (1994)a
256
Chapter 7
Flavor Limit
CL
(5 eV)
0.16 MeV
23.8 MeV
Reference
Nonstandard Neutrinos
257
95% CL mass limit of 23.8 MeV on the basis of 25 events of the form
5 and 5 where 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 e the
observed neutrinos from SN 1987A gave me <
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
mass down to the cosmologically interesting 30 eV range (Sect. 11.6)!
For Dirac neutrinos there is another essentially kinematical constraint from the SN 1987A neutrino observations. The sterile Dirac
components can be produced in scattering processes by helicity ips.
In a supernova core this eect leads to an anomalous energy drain,
limiting a Dirac mass to be less than a few 10 keV (Sect. 13.8.1).
7.1.4
258
Chapter 7
3
11
mi ,
(7.2)
i=1
mi
,
i=1 93 eV
(7.3)
(7.4)
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 me + m + m 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 avor.
If neutrinos were unstable and if they decayed so early that their
decay products were suciently redshifted by the expansion of the universe, the cosmological mass bound can be violated without running
into direct conict with observations. The excluded range of masses
Nonstandard Neutrinos
259
Note that the corresponding limits discussed in the book by Kolb and Turner
(1990) are somewhat less restrictive because their treatment does not seem to be
entirely self-consistent (G. Gelmini, private communication).
260
Chapter 7
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 bigbang nucleosynthesis (BBN) argument has been used to constrain the
number of light neutrino families to N <
3.4 (Yang et al. 1984; Olive
et al. 1990). Even though the measured Z decay width has established
N = 3 (Particle Data Group 1994) the BBN bound remains of interest
as a mass limit because massive neutrinos contribute more than a massless one to the expansion rate at BBN. For a lifetime exceeding about
<
100 s this argument excludes 500 keV <
m 35 MeV (Kolb et al. 1991;
Dolgov and Rothstein 1993; Kawasaki et al. 1994), with even more restrictive 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
(Sect. 15.7) as a specic model for the neutrino decay. Including the
energy density of the scalar they nd even more restrictive limits
which exclude the region between the dashed lines in Fig. 7.2.
7.2
7.2.1
One of the most mysterious features of the particle zoo is the threefold
repetition of families (or avors) shown in Fig. 7.1. The fermions in
each column have been arranged in a sequence of increasing mass which
appears to be the only signicant dierence between them. There is no
indication for higher sequential families; the masses of their neutrinos
would have to exceed 12 mZ = 46 GeV according to the CERN and SLAC
measurements of the Z decay width (Particle Data Group 1994). If
the origin of masses is indeed the interaction with the vacuum Higgs
eld, the only dierence between the fermions of a given column in
Fig. 7.1 is their Yukawa coupling to .
If the only dierence between, say, an electron and a muon is the
vacuum refraction, any superposition between them is an equally legitimate charged lepton except for the practical diculty of preparing it
experimentally. When such a mixed state propagates, the two components acquire dierent phases along the beam exactly like two photon
helicities in an optically active medium, leading to a rotation of the
plane of polarization. Of course, now this polarization is understood
in the abstract avor space rather than in coordinate space.
Nonstandard Neutrinos
261
In three-dimensional avor space one is free to choose any superposition of states as a basis. It is convenient and common practice to use
the mass eigenstates (vacuum propagation eigenstates) for each column
of Fig. 7.1. Thus by denition the electron is the charged lepton with
the smallest mass eigenvalue, the muon the second, and the tau the
heaviest, and similarly for the quarks.
All fermions interact by virtue of the weak force and thus couple to
the W and Z gauge bosons, the quarks and charged leptons in addition couple to photons, while only the quarks interact by the strong
force and thus couple to gluons. The W (charged current) interaction has the important property of changing, for example, a charged
lepton into a neutrino as in the reaction p + e n + e (Fig. 7.3). If
the initial charged lepton was an electron (the lightest charged lepton
mass eigenstate), the outgoing neutrino state is dened to be an electron neutrino or e which in general will be a certain superposition of
neutrino mass eigenstates.
d
s
cos C
sin C
sin C
cos C
)(
d
.
s
(7.5)
262
Chapter 7
Including the third generation, the mixing is induced by the threedimensional Cabbibo-Kobayashi-Maskawa (CKM) matrix V . After removing all unphysical phases by an appropriate redenition of the quark
elds this unitary matrix is given in terms of four signicant parameters.
The Particle Data Group (1994) recommends a standard parametrization in terms of three two-family mixing angles ij < /2 and one phase
0 < 2. With Cij cos ij and Sij sin ij this standard form is
(Fritzsch and Plankl 1987)
1
0
V = 0 C23
0 S23
0
C13
S23
0
S13 ei
C23
C12 C13
S12
S12 S23 S13 ei
S12
1
S23
0 S13 ei
C12
1
0 S12
0
C13
0
S12 C13
C12 C23 S12 S23 S13 ei
C12 S23 C23 S12 S13 ei
S12
C12
0
0
1
S13 ei
C13 S23
C13 C23
S13 ei
S23 .
1
(7.6)
Experimentally one has the 90% CL ranges (Particle Data Group 1994)
0.218 < S12 < 0.224,
0.032 < S23 < 0.048,
0.002 < S13 < 0.005.
(7.7)
Ui i ,
(7.8)
i=1
where the unitary matrix U plays the role of the CKM matrix. Unless
otherwise stated 1 will always refer to the dominant mass admixture
of e 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 dierent family. This phenomenon 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 example, muons decay only by the avor-conserving mode e e
(Fig. 7.4); the experimental upper limits on the branching ratios for
e and e e+ e are 51011 and 1.01012 , respectively.
In the absence of neutrino masses and mixing the individual lepton avor 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.
264
Chapter 7
= |Ueh |2
G2F
m5 (mh )
3 (4)3 h
(7.9)
(7.10)
The one- and two-photon decay modes arise in the standard model
with mixed neutrinos from the amplitudes shown in Fig. 7.7. Turn rst
to the one-photon decay i j 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.
(7.11)
m2i m2j
mi
)3
1
= 5.308 s
e
B
)2
3
m
m3eV ,
(7.12)
where 2e |ij |2 + |ij |2 , meV mi /eV, and m (m2i m2j )/m2i .
An explicit evaluation of the one-photon amplitude of Fig. 7.7 yields
for Dirac neutrinos (Pal and Wolfenstein 1982)
D
e 2 GF
ij
(m
m
)
Uj Ui f (r ).
(7.13)
=
i
j
2
(4)
D
=e,,
ij
D
M
For Majorana neutrinos one has instead M
ij = 2ij and ij = 0 or
M
M
D
ij = 0 and ij = 2ij , depending on the relative CP phase of i and j .
In Eq. (7.13) r (m /mW )2 where the charged-lepton masses are
me = 0.511 MeV, m = 105.7 MeV, and m = 1.784 GeV while the
W gauge boson mass is mW = 80.2 GeV. Thus for all charged leptons
r 1; in this limit
f (r ) 32 + 34 r .
(7.14)
If one inserts the leading term 32 into the sum in Eq. (7.13) one nds
that its contribution vanishes because the unitarity of U implies that its
rows or columns represent orthogonal vectors. Because the rst nonzero
266
Chapter 7
3GF me
=
(mi mj )
2 (4)2
23
= 3.9610
mi mj
1 eV
m
mW
)2
=e,,
Uj Ui
=e,,
Uj Ui
m
m
m
m
)2
)2
(7.15)
These small numbers imply that neutrino radiative decays are exceedingly 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 electromagnetic moments. For D
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,
D
2 G F me
6
ii
=
mi = 3.201019 meV ,
(7.16)
B
(4)2
much larger than the transition moments because it is not GIM suppressed.
The two-photon decay rate i j is of higher order and thus
may be expected to be smaller by a factor of /4. However, it is
not GIM suppressed so that it is of interest for a certain range of neutrino masses (Nieves 1983; Ghosh 1984). Essentially, the result involves another factor /4 relative to the one-photon rate, and f (r )
in Eq. (7.13) is replaced by (mi /m )2 .
As an example consider the dierent decay modes for 3 1 ,
assuming that m3 m1 and that the mixing angles are small so that
3 and 1 e . Then one has explicitly
(m3 ),
(
)
m 4
27
3 1 e+ e ,
G2 m 5
1
,
3 1 ,
|Ue3 |2 F 33 8 4 mW
3 (4)
(
) (
)
1
2 m3 4
, 3 1 ,
180 4
me
(7.17)
where (mh ) was given in Eq. (7.10) and shown in Fig. 7.6. The
decay dominates in a small range of m3 just below 2me .
Nonstandard Neutrinos
267
7.3
7.3.1
When Wolfgang Pauli in 1930 rst postulated the existence of neutrinos he speculated that they might interact like a magnetic dipole of a
certain moment . If that were the case they could be measured by
their ionizing power when they move through a medium; this ionizing
power was rst calculated by Bethe (1935). Nahmias (1935) measured
the event rates in a Geiger-M
uller counter in the presence and absence of a radioactive source and interpreted his null result as a limit
4
<
210 B (Bohr magneton B = e/2me ) on the neutrino dipole
moment. He concluded that since this limit is already smaller than a
nuclear magneton, it seems probable that the neutrino has no moment
at all. Subsequent attempts to measure ever smaller neutrino dipole
moments have consistently failed.
The main dierence between then and now is the advanced theoretical understanding of neutrino interactions in the context of the
standard model of electroweak gauge interactions. A magnetic dipole
interaction couples l.h. with r.h. states so that the latter would not
be strictly sterile. This would be in conict with the standard model
where neutrinos interact only by their l.h. coupling to W and Z gauge
bosons. Thus neutrino dipole moments must vanish identically because
weak interactions violate parity maximally.
This picture changes when neutrinos have masses because even the
r.h. components of a Dirac neutrino are then not strictly sterile as
they couple to the Higgs eldor else they would not have a mass.
Indeed, an explicit calculation in the standard model with neutrino
masses gave a magnetic dipole moment = 3.201019 B (m /eV)
(Eq. 7.16). If neutrinos mix, they also obtain transition magnetic and
electric moments. However, they are even smaller because of the GIM
suppression eectsee Eq. (7.15).
268
Chapter 7
Much larger values would obtain with direct r.h. neutrino interactions. 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 dierence between the l.h. and r.h. gauge
bosons. For a neutrino (avor = e, or ) the dipole moment in
such models is (Kim 1976; Marciano and Sanda 1977; Beg, Marciano,
and Ruderman 1978)
eGF
= 2
2 2
m2W1
1 2
mW2
)
3
4
sin 2 + m
m2W1
1+ 2
m W2
)]
,
(7.18)
Single-Photon Coupling
There are many possible extensions of the standard model which would
give sizeable neutrino dipole and transition moments by some novel r.h.
interaction. For the purposes of this book the underlying new physics
is of no concern; all we need is a generic representation of its observable
eects in terms of neutrino electromagnetic form factors.
The most general interaction structure of a fermion eld with the
electromagnetic eld can be expressed as an eective Lagrangian
Lint = F1 A G1 5 F
12 (F2 + G2 5 )F ,
(7.19)
Nonstandard Neutrinos
269
F1 (Q2 )
r = 6
e Q2 Q2 =0
2
(7.20)
Babu and Mohapatra 1990; Babu and Volkas 1992; Takasugi and Tanaka 1992;
Foot, Lew, and Volkas 1993; Foot 1994. References to earlier works are given in
these papers.
270
Chapter 7
H = i + m ( + i5 )(i E + B) ,
(7.21)
,
0
I 0
0 I
0
,
(7.22)
For recent discussions of these matters see Lucio, Rosado, and Zepeda (1985),
Auriemma, Srivastava, and Widom (1987), Degrassi, Sirlin, and Marciano (1989),
Musolf and Holstein (1991), and Gongora-T. and Stuart (1992).
Nonstandard Neutrinos
271
(7.23)
Two-Photon Coupling
Discussions of neutrino electromagnetic form factors are usually restricted to the eective neutrino coupling to an electromagnetic wave
or static eld. However, a two-photon coupling of neutrinos is also possible and of some interest. Historically, it was thought for some time
that the process could be of great importance for the emission of neutrinos from stars until it was shown by Gell-Mann (1961)
that the amplitude for this process vanishes identically if neutrinos
have only l.h. local interactions with electrons. Several authors discussed the process when neutrinos are massive, or when they
have more general interaction structures (Halprin 1975; Fischbach et al.
1976, 1977; Natale, Pleitez, and Tacla 1987; Gregores et al. 1995). However, there does not seem to be a plausible scenario where this process
would be of serious astrophysical interest.
In the standard model, there is an eective two-photon coupling
to neutrinos because the interaction is not local; rather, it is mediated by nite-mass gauge bosons. Early calculations of the eective
45
For example Nieves (1982), Kayser (1982), Shrock (1982), and Li and Wilczek
(1982).
272
Chapter 7
7.4
Electromagnetic Processes
G2 m e
T
d
= F
(CV + CA )2 + (CV CA )2 1
dT
2
E
]
+ (CA2
CV2 )
)2
]
me T
1
1
+ 2
,
2
E
T
E
(7.24)
Nonstandard Neutrinos
273
Fig. 7.8. Important processes involving direct neutrino electromagnetic couplings, notably magnetic dipole moments and magnetic or electric transition
moments.
1
2
2 r /GF , i.e. there would be a small correction to the overall
3
cross section. A dipole moment, on the other side, modies the energy
spectrum of the recoil electrons, notably at low energies, because of its
forward-peaked nature from the Coulomb divergence.
If neutrinos had electric dipole moments, or electric or magnetic
transition moments, these quantities would also contribute to the scattering cross section. Therefore, the quantity measured in electron recoil
experiments involving relativistic neutrinos is (e.g. Raelt 1989)
2 =
|ij ij |2 ,
(7.25)
j=e , ,
274
Chapter 7
A related process to scattering by photon exchange is the spinprecession in a macroscopic magnetic or electric eld into r.h. states of
the same or another avor, i.e. electromagnetic oscillation into wronghelicity states. Such eects will be studied in Sect. 8.4 in the general
context of neutrino oscillations. In principle, this process can be important at modifying the measurable l.h. solar neutrino ux as discussed
in Sect. 10.7 in the context of the solar neutrino problem. Spin oscillations can also be important in supernovae where strong magnetic elds
exist, although a detailed understanding remains elusive at the present
time (Sect. 11.4).
The most interesting process caused by dipole moments is the photon decay into neutrino pairs, , which is enabled in media where
the photon dispersion relation is such that 2 k2 > 0. It is the most
interesting process because it occurs even in the absence of dipole moments due to a medium-induced neutrino-photon coupling (Chapter 6).
This is the dominant standard neutrino emission process from stars for
a wide range of temperatures and densities. Details of both the standard and the dipole-induced plasma process depend on complicated
ne points of the photon dispersion relation in media that were taken
up in Chapter 6.
The salient features, however, can be understood in an approximation where photons in a medium (transverse plasmons) are treated as
particles with an eective mass equal to the plasma frequency P which
in a nonrelativistic medium is P2 = 4ne /me with the electron density ne and electron mass me . The decay rate of these electromagnetic
excitation in their own rest frame is then
=
P3
24
with
2 =
|ij |2 + |ij |2 ,
(7.26)
i,j
while in the frame of the medium where the photon has the energy a
Lorentz factor P / must be included. The sum includes all nal-state
neutrino avors with mi P ; otherwise phase-space modications
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 eects between magnetic and electric
dipole amplitudes occur.
Transition moments would allow for the radiative decay i j .
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
m3i
8
with
2 =
|ij |2 + |ij |2 .
(7.27)
7.5
7.5.1
276
Chapter 7
(7.28)
(7.29)
(7.30)
(Cooper-Sarkar et al. 1992). Of course, in this range the electromagnetic cross section would far exceed a typical weak interaction one.
Nonstandard Neutrinos
7.5.2
277
Neutrino scattering by photon exchange can also be important in astrophysical settings, notably if neutrinos are Dirac particles. The magnetic
or electric dipole coupling is such that it ips the helicity of relativistic
neutrinos, i.e. the nal state is r.h. for an initial l.h. neutrino. This
spin ip is of no importance in experiments where the electron recoil is
measured, but it can have dramatic consequences in supernovae where
l.h. neutrinos are trapped by the standard weak interactions. The spinip scattering by a electromagnetic dipole interaction would produce
wrong-helicity states that could freely escape unless they scattered
again electromagnetically. The SN 1987A neutrino signal indicates that
this anomalous cooling channel cannot have been overly eective, yield12
ing a constraint of around <
310 B on all Dirac diagonal or
transition moments in the sense of Eq. (7.25); see Sect. 13.8.3 for a
more detailed discussion.
One should keep in mind, however, that for dipole moments in this
range the spin precession in the strong macroscopic magnetic elds that
are believed to exist in and near SN cores could also cause signicant
left-right transitions. Notably, the back conversion of r.h. neutrinos
could cause a transfer of energy between widely separated regions of the
SN core, and might even help at the explosion (Sect. 13.8.3). The role
of relatively large neutrino dipole moments in SN physics has not been
elaborated in enough depth to arrive at reliable regions of parameters
that are ruled out or ruled in by SN physics and the SN 1987A neutrino
signal.
7.5.3
278
Chapter 7
(7.31)
The search for radiative decays of reactor, beam, solar, supernova, and
cosmic neutrinos will be discussed at length in Chapter 12. For electron
neutrinos, the eective transition moment in the sense of Eq. (7.27) will
be found to be limited by
1
2
0.910 B (eV/m )
0.5105 (eV/m )2
B
<
1.510 B (eV/m )2
10
2.3
310
B (eV/m )
Reactors,
Sun,
SN 1987A,
Cosmic background,
(7.32)
Nonstandard Neutrinos
279
The last and most interesting constraint arises from the energy-loss
argument applied to globular cluster stars. The neutrino emissivity by
the plasma process would be too large unless
12
<
310 B .
(7.33)
Chapter 8
Neutrino Oscillations
The phenomenon of neutrino oscillations in vacuum and in media as
well as in magnetic elds is studied. Experimental constraints on neutrino mixing parameters are reviewed.
8.1
Introduction
280
Neutrino Oscillations
281
282
Chapter 8
oscillations, except that here the two helicity components rather than
the avor components get transformed into each other. Again, this is
a standard eect familiar from the behavior of electrons in magnetic
elds. In astrophysical bodies large magnetic elds exist, especially in
supernovae, so that magnetic helicity oscillations are potentially interesting. However, much larger magnetic dipole moments are required
than are predicted for standard massive neutrinos. Thus, avor oscillations have rightly received far more attention.
Presently I will develop the theoretical tools for neutrino oscillations, and summarize the current experimental situation. In Chapter 9
I will discuss the more complicated phenomena that obtain when oscillating neutrinos are trapped in a supernova core. The story of solar
neutrinos (Chapter 10) is inextricably intertwined with that of neutrino oscillations, especially of the MSW variety. Finally, oscillations
may also play a prominent role for supernova neutrinos and the interpretation of the SN 1987A signal (Chapter 11).
8.2
Vacuum Oscillations
8.2.1
In order to derive a formal equation for the oscillation of mixed neutrinos I begin with the equation of motion of a Dirac spinor i which
describes the neutrino mass eigenstate i. It obeys the Dirac and thus
the Klein-Gordon equation (t2 2 + m2i ) i = 0. One may readily
combine all mass eigenstates in a single equation
(t2 2 + M 2 ) = 0,
where
m21
2
M 0
0
0
m22
0
0
m23
(8.1)
1
and 2 .
3
(8.2)
Eq. (8.1) may be written in any desired avor basis, notably in the
basis of weak-interaction eigenstates to which one may transform by
virtue of Eq. (7.8),
1
e
= U 2 .
3
(8.3)
Neutrino Oscillations
283
As usual one expands the neutrino elds in plane waves of the form
(t, x) = k (t) eikx for which Eq. (8.1) is
(t2 + k2 + M 2 ) k (t) = 0 .
(8.4)
it k = k k
M2
where k k +
.
2k
(8.5)
(8.6)
iz = K
M2
.
where K
2
(8.7)
This equation describes the spatial variation of a neutrino beam propagating in the positive z-direction with a xed energy .
Ultimately one is not interested in amplitudes but in the observable
probabilities | |2 = with = e, , or . One may derive an
284
Chapter 8
(8.8)
(8.9)
(8.10)
(8.11)
Two-Flavor Oscillations
Neutrino Oscillations
285
(8.12)
with the mixing angle , the 22 unit matrix I, and the Pauli matrix47
2 . The mass matrix may be written in the form
M 2 /2 = b0 12 B ,
(8.13)
sin 2
2
B=
(8.14)
0 ,
osc
cos 2
a vector which is tilted with regard to the 3-axis by twice the mixing
angle (Fig. 8.2). Further,
4
osc 2
(8.15)
m2 m21
is the oscillation length. Its meaning will presently become clear.
In this representation it is straightforward to work out the spatial
behavior of a stationary neutrino beam. From K = b0 + 21 B
one nds
[
(
)
(
)(
)]
z
z
cos 2 sin 2
W = ei(b0 )z cos
i sin
.
sin 2
cos 2
osc
osc
(8.16)
Assuming that the oscillations are among the rst two families, the
appearance probability for a and the e survival probability are for
an initial e
prob (e ) = |We |2 = sin2 (2) sin2 (z/osc ),
prob (e e ) = |Wee |2 = 1 prob (e ).
(8.17)
0
1
)
(
)
(
)
1
0 i
1 0
, 2 =
, and 3 =
.
0
i 0
0 1
286
Chapter 8
(8.18)
(8.19)
Neutrino Oscillations
287
8.2.3
If the neutrino source region is not point-like relative to the oscillation length, one has to average the appearance or survival probabilities
accordingly. If the source locations z0 are distributed according to a
normalized function f (z0 ) the appearance probability is
prob (e ) = sin2 2
(z z0 )
.
osc
(8.20)
2
2
For example, consider a Gaussian distribution f (z0 ) = ez0 /2s /s 2
of size s for which
prob (e ) =
1
2
sin2 2 1 e2
2 (s/
2
osc )
cos(2z/osc ) .(8.21)
For s = 15 osc this result is shown in Fig. 8.3. For s = 0 Eq. (8.21) is
identical with Eq. (8.17) while for s osc it is 12 sin2 2 which reects
that the beam is an incoherent mixture: the relative phases between
dierent avor components have been averaged to zero.
No source is exactly monochromatic; usually the neutrino energies
are broadly distributed. With a point source and a normalized distribution g() one nds
(m22 m21 ) z
.
(8.22)
4
As an example let g() such
that = 2/osc follows a Gaussian
(0 )2 /2 2
distribution e
/ 2 of width and with 0 = 2/0 . Then
prob (e ) = sin2 2
prob (e ) =
1
2
d g() sin2
sin2 2 1 e
2 z 2 /2
cos(2z/0 )
(8.23)
1
which is shown in Fig. 8.4 for = 10
0 . For = 0 Eq. (8.23) reproduces
Eq. (8.17) while for z 1 it approaches 12 sin2 2.
288
Chapter 8
Neutrino Oscillations
8.2.4
289
Because neutrino masses must be very small the oscillation length involves macroscopic scales. Numerically, it is
osc = 2.48 m
E 1 eV2
.
1 MeV m2
(8.24)
Fig. 8.5. Experimental limits on neutrino masses and mixing angles. Reactors, e disappearance: (a) Bugey 4 (Achkar et al. 1995), superseding the
G
osgen limits (Zacek et al. 1986); (b) Kurchatov Institute (Vidyakin et al.
1987, 1990, 1991). Accelerator experiments: (c) BNL Experiment 776, wideband beam, e and e appearance (Borodovsky et al. 1992). (d) BNL Experiment 734, measurement of e / ratio (Ahrens et al. 1985). (e) Fermilab
Experiment 531, appearance (Ushida et al. 1986); similar constraints were
reported by the CHARM II Collaboration (1993). (f) CDHS Experiment,
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-avor oscillations 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 expected to see oscillations in the e or channel if the
neutrino mixing angles are comparably small. It is very encouraging
that the NOMAD and CHORUS experiments which are currently taking 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 m in the 10 eV range and the small
mixing angles probed. Other future but less advanced projects for terrestrial 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 e oscillations
(Athanassopoulos et al. 1995). If this interpretation is correct, the
corresponding m2 would exceed about 1 eV2 , while sin2 2 would be a
few 103 . 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.
8.2.5
Atmospheric Neutrinos
Besides reactors and accelerators, one may also use atmospheric neutrinos as a source to search for oscillations. Primary cosmic ray protons produce hadronic showers when interacting with atmospheric nuclei (A). Neutrinos are subsequently produced according to the simple
scheme
p + A n + /K + . . .
/K + ( ) + ( )
+ ( ) e+ (e ) + e ( e ) + ( ).
(8.25)
Therefore, one expects twice as many s as e s, and equally many
neutrinos as antineutrinos of both avors. At a detector, the neutrino ux is approximately isotropic except at energies below about
1 GeV where geomagnetic eects become important. Because the neutrinos come from anywhere in the atmosphere, from directly overhead
or from as far as the antipodes, oscillation lengths between about 10
Neutrino Oscillations
291
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 e / anomaly at Kamiokande (Fukuda et al. 1994);
the star marks the best-t value for the mixing parameters. The hatched
areas are excluded by: (b) e / ratio at Frejus (Frejus Collaboration 1990,
1995; Daum 1994). (c) Absolute rate and (d) stopping fraction of upward going 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 normalization of the ux must be determined by calculations and thus is
probably uncertain to within about 30% while the e / avor ratio
is likely known to within, say, 5%.
Several underground proton decay experiments have reported measurements of atmospheric neutrinos. The Frejus detector (an iron
calorimeter) saw the expected e / avor ratio and thereby excluded
the range of masses and mixing angles marked b in Fig. 8.6 for e -
and - oscillations (Frejus Collaboration 1990, 1995; Daum 1994).
Instead of measuring the neutrinos directly one may also study the
ux of secondary muons produced by interactions in the rock surround48
The eect of matter must be included for e - atmospheric neutrino oscillations. For a recent detailed analysis see Akhmedov, Lipari, and Lusignoli (1993).
292
Chapter 8
ing the detector. (Electrons from e interactions range out much faster
in the rock and so one expects mostly muons from s.) This method is
sensitive to the high-energy spectral regime of the atmospheric ux.
Moreover, one may select upward going muons which are produced from
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 decit of atmospheric
s relative to e s, a nding usually expressed in terms of a ratio of
ratios, i.e. the measured over the expected ratio of e-like over -like
events (Fig. 8.7). While this procedure is justied because it is largely
free of the uncertain absolute ux normalization, one must be careful
at interpreting the signicance of the ux decit. The error of a measured ratio does not follow a Gaussian distribution; a representation
like Fig. 8.7 tends to overemphasize the signicance of the discrepancy
(Fogli and Lisi 1995).
Neutrino Oscillations
293
Apparently, the anomaly observed at the Kamiokande water Cherenkov detector (Hirata et al. 1992; Fukuda et al. 1994) can be explained in terms of oscillations for neutrino parameters in the shaded
area in Fig. 8.6; the best-t value is indicated by a star. These results
are a combined t for the sub-GeV and multi-GeV data as published
by the Kamiokande collaboration (Fukuda et al. 1994). The oscillation hypothesis appears to be buttressed by a zenith-angle variation of
the eect observed in Kamiokandes multi-GeV data sample although
the claimed signicance of this eect has been critiqued, e.g. by Fogli
and Lisi (1995) and by Saltzberg (1995).
The required large mixing angle as well as the exclusion regions
of the other experiments make it appear dubious that the anomaly is
caused by oscillations. Still, it is a serious eect that cannot be blamed
easily on problems with the Kamiokande detector. Also, the reliability
of some of the exclusion areas in Fig. 8.6 may be called into question,
notably because of their dependence on absolute ux normalizations.
The intuition against a large - mixing angle may be misguided.
In the future, it will be possible to test the relevant regime of mixing
parameters in long-baseline laboratory experiments (e.g. Schneps 1995).
At the time of this writing, the possibility that the atmospheric neutrino
anomaly may be revealing neutrino oscillations remains a lively-debated
possibility.
8.3
8.3.1
Oscillations in Media
Dispersion Relation for Mixed Neutrinos
294
Chapter 8
3Ye 1
G
n
F B
0
0
Ye 1
0
0
Ye 1
m21
k 2 U 0
0
0
e
0 U = 0,
m23
0
m22
0
(8.26)
(8.27)
2
where it is easy to read the matrix Me
from Eq. (8.26). Because to
2
lowest order = k one may equally use Me
/2 in order to derive the
dispersion relation, depending on whether one wishes to write as a
function of k or vice versa.
For two-avor mixing between e and or (mixing angle 0 ) the
eective mass matrix may be written in the same form as in vacuum
2
Me
/2 = b0 12 B ,
(8.28)
2 GF nB (Ye 12 ) and
sin 2
sin 20
0
m22 m21
2
0 2 GF n e 0 .
B=
0 =
osc
2
cos 20
cos 2
1
(8.29)
This equation denes implicitly the mixing angle as well as the oscillation length osc in the medium in terms of the masses, the vacuum
mixing angle 0 , 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 20 = 0.15.
296
Chapter 8
Explicitly one nds for the mixing angle and the oscillation length
in the medium the following expressions,
sin 20
,
cos 20
sin 20
sin 2 =
,
2
[sin 20 + (cos 20 )2 ]1/2
tan 2 =
(8.30)
where
Ye
eV2
2 GF ne 2
7
=
1.5310
,
m22 m21
g cm3 MeV m22 m21
(8.31)
and
osc =
4
sin 2
.
2
m1 sin 20
m22
(8.32)
For m2 > m1 these functions are shown in Fig. 8.8; they exhibit a
resonance for cos 20 = .
The dispersion relation has two branches which in vacuum correspond to 1,2 = (k 2 m21,2 )1/2 . In the relativistic limit they are
1,2 k =
m21 + m22
+ 2 GF nB (Ye 12 )
4k
]1/2
m2 m21 [ 2
2
sin 20 + (cos 20 )2
,
4k
(8.33)
a result schematically shown in Fig. 8.8. The resonance of the mixing 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 eect changes sign for antineutrinos, a resonance occurs between e and if m2 > m1 , while none occurs between
e and . If the mass hierarchy is the other way round, a resonance
occurs for e and , but not for e and .
8.3.2
Neutrino Oscillations
297
when the quantity of Eq. (8.31) is much smaller than unity. In the
opposite limit one nds for the mixing angle in the medium
m2 m21
sin 2 = 2
sin 20 .
2 GF ne 2
(8.34)
2
g cm3
= 1.63104 km
,
Ye
2 GF ne
(8.35)
W = S exp i
K(z ) dz
(8.36)
298
Chapter 8
(8.37)
(8.38)
Neutrino Oscillations
299
If the neutrino crosses a density region such that a resonance occurs, this part of the trajectory yields the most restrictive adiabaticity
requirement. On resonance = cos 20 and sin 2 = 1. In this case one
denes an adiabaticity parameter
m22 m21 sin 20 tan 20
,
(8.39)
2
| ln ne |res
where the denominator is to be evaluated at the resonance point. The
adiabatic condition is 1. It establishes a relationship between vacuum mixing angles and neutrino masses for which resonant oscillations
occur.
8.3.4
1
2
+ ( 21 p) cos 20 cos 2,
(8.40)
300
Chapter 8
the other (Fig. 8.8, lowest panel) when it moves across the resonant
density region.
A linear density prole near the resonance region, which is always
a rst approximation, yields the Landau-Zener probability
p = e/2
(8.41)
which was rst derived in 1932 for atomic level crossings. The adiabaticity parameter was dened in Eq. (8.39); in the adiabatic limit
1 one recovers p = 0.
For a variety of other density proles and without the assumption
of a small mixing angle one nds a result of the form
e(/2) F e(/2) F
p=
,
1 e(/2) F
(8.42)
(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 specic forms for ne (r) for which analytic results exist. Recently, Guzzo, Bellandi, and Aquino (1994) used a somewhat
dierent 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
(
p=
1
1 +
)2
sin2 (0 ) +
[
]
2
2
2
cos
(
)
+
cos
(
+
)
,
0
0
(1 + )2
(8.44)
Neutrino Oscillations
8.3.5
301
1
6
m2 MeV
.
meV2 E
(8.45)
meV2 E
.
m2 MeV
(8.46)
cos 20 c
.
[(cos 20 c )2 + sin2 20 ]1/2
(8.47)
302
Chapter 8
where they are not. The dotted line marks c = cos 20 and thus indicates for which neutrino parameters a resonance occurs on the way out
of the Sun.
Fig. 8.9. The MSW triangle for the simplied solar model discussed in the
text with E = 1 MeV.
Fig. 8.10. The MSW bathtub for the simplied solar model discussed in the
text with m2 = 3105 eV2 and sin2 20 = 0.01.
Neutrino Oscillations
303
304
8.4
8.4.1
Chapter 8
(8.49)
Neutrino Oscillations
305
system, however, involves a time-dilation factor 1 so that the Hamiltonian for the spin evolution in the laboratory system is H = 1 B
or explicitly
[
)
(B v
) + E v .
H = (m/)(B v
v+v
(8.50)
1 cos
H = B +i
e sin
ei sin
,
1 cos
(8.51)
H = BT
0 1
.
1 0
(8.52)
Therefore, if one begins with left-handed neutrinos a complete precession into right-handed ones will always occur (assuming a sucient
path length), independently of the direction between the laboratory
magnetic eld and the neutrino direction of motion (Fujikawa and
Shrock 1980).
If there is only an electric eld in the laboratory frame, magnetic
oscillations will nevertheless occur because Eq. (8.50) implies that the
neutrino sees an eective magnetic eld BT = E v (Okun 1986). The
precession thus takes place in the plane of E and v.
Contrary to avor oscillations, the oscillation length does not depend on the energy. Therefore, a broad energy spectrum would not
cause a depolarization of the neutrino ux from an astrophysical source
(Sun, supernovae). Rather, neutrinos of all energies would spin-precess
in step with each other.
Because the precession frequency is 2BT the neutrinos will have
reversed their spin after a distance
1
2 osc
1010 B 1 G
= 5.361013 cm
,
2BT
BT
(8.53)
306
Chapter 8
substantial reduction of the left-handed solar neutrino ux then requires magnetic dipole moments of order 1010 B . Typical galactic
magnetic elds are of order 106 G with coherence scales of order kpc.
Therefore, neutrinos from a galactic source such as a supernova could
be ipped before reaching Earth if their dipole moment is of order
1012 B . These values are in the neighborhood of what is experimentally and observationally allowed so that there remains interest in the
possibility that magnetic spin oscillations could have a signicant impact on the neutrino uxes from the Sun or from supernovae.
8.4.2
H=
0
BT
BT
,
2GF (ne 12 nn )
(8.54)
GF (ne 21 nn )
1010 B (Ye 12 Yn )
= 66.6 kG
.
g cm3
2
(8.55)
Here, is the mass density and Ye,n the electron and neutron number
per baryon. If BL BT the precession is around a direction close to
the direction of motion, and so the spin reversal will always be far from
complete. Hence, the presence of the medium suppresses magnetic spin
oscillations (Voloshin, Vysotski, and Okun 1986).
8.4.3
Spin-Flavor Precession
Neutrino Oscillations
307
it
m2e /2p
BT
BT
m2 /2p
)(
e
.
(8.56)
m c2
e
s
it = m 2
e
0
BT
m s2
m c2
BT
0
0
BT
m c2
m s2
BT
e
0
,
m s2 e
m c2
(8.57)
m c2 + Ve
m s2
0
BT
m s2
m c2 + V
BT
0
0
BT
m c2 Ve
m s2
BT
. (8.58)
m s2
m c2 V
If there are density gradients (solar neutrinos!) one may have resonant
magnetic conversions between, say, e and where the barrier to spinprecessions caused by the mass dierence is compensated by matter
eects, i.e. one may have resonant spin-avor oscillations (Akhmedov
1988a,b; Barbieri and Fiorentini 1988; Lim and Marciano 1988).
For Dirac neutrinos, transitions to antineutrinos are not possible
and so one needs to consider oscillations separately in the four-level
308
Chapter 8
R
system (eL , L , eR , R ) and ( Le , L , R
e , ). Moreover, one may have
both diagonal and transition magnetic moments so that for neutrinos
the r.h.s. of the equation of motion is
m c2 + Ve
m s2
ee BT
e BT
m s2
m c2 + V
e BT
BT
ee BT
e BT
m c2
m s2
eL
e BT
BT L
m s2 eR
R
m c2
(8.59)
BL
H=
BT ei
BT ei
,
BL
(8.60)
Neutrino Oscillations
309
U=
ei/2
0
ei/2
(8.61)
BL
H =
BT
BT
BL
1 0
.
0 1
(8.62)
Put another way, the eective longitudinal magnetic eld in the rotating
frame is BL = BL 12 /
while BT remains unchanged. |BL | can be less
or larger than |BL | so that the transition rate between the two levels
can be increased or decreased by a twist. BL may even be cancelled
entirely, enabling resonant oscillations.
For a systematic discussion of the full four-level spin-avor problem
see, for example, Akhmedov, Petcov, and Smirnov (1993).
Chapter 9
Oscillations
of Trapped Neutrinos
In a supernova (SN) core or in the early universe neutrinos scatter on
the background medium and on each other. Therefore, oscillations of
mixed neutrinos are frequently interrupted, leading to avor equilibrium if there is enough time. A Boltzmann-type kinetic equation is
derived that accounts simultaneously for neutrino oscillations and collisions for arbitrary neutrino degeneracy. It is applied to a SN core,
yielding an estimate of the time scale to reach avor equilibrium. On
the basis of the SN 1987A neutrino observations a limit on the mixing
of sequential neutrinos with a hypothetical sterile avor is derived.
9.1
Introduction
311
neutrino avor (say e ) scatters with a rate while the other (say )
does not. An initial e will begin to oscillate into . The probability
for nding it in one of the two avors evolves as previously discussed
and as shown in Fig. 9.1 (dotted line). However, in each collision the
momentum of the e component of the superposition is changed, while
the component remains unaected. Thus, after the collision the two
avors 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 e to develop a new coherent 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 e s and s.
This decoherence eect is even more obvious when one includes the
possibility of e absorption and production by charged-current reactions
e n pe. Because of oscillations an initial e is subsequently found
to be a with an average probability of 21 sin2 2 (mixing angle )
and as such cannot be absorbed, or only by the reaction n p
if it has enough energy. The continuous emission and absorption of
e s spins o a with an average probability of 12 sin2 2 in each
collision! Chemical relaxation of the neutrino avors will occur with
an approximate rate 21 sin2 2 where is a typical weak interaction
rate for the ambient physical conditions. An initial e population
turns into an equal mixture of e s and 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
A certain damping of avor oscillations occurs even without collisions when the neutrinos are not monochromatic because then dierent
modes oscillate with dierent frequencies. This dephasing eect was
shown in Fig. 8.4 and is repeated in Fig. 9.1 (dashed line). While the
dephasing washes out the oscillation pattern, it does not lead to avor
equilibrium: the probability for e ends at a constant of 1 12 sin2 2.
For the rest of this chapter damping of oscillations never refers to
this relatively trivial dephasing eect.
The interplay of collisions and oscillations leads to avor equilibrium
between mixed neutrinos. In a SN core the concentration of electron
lepton number is initially large so that the e form a degenerate Fermi
sea. The other avors and are characterized by a thermal distribution at zero chemical potential. However, if they mix with e they
will achieve the same large chemical potential. In a SN core heat and
lepton number are transported mostly by neutrinos; the eciency of
these processes depends crucially on the degree of neutrino degeneracy
for each avor. Therefore, it is of great interest to determine the time
it takes a non-e avor to equilibrate with e under the assumption
of mixing (Maalampi and Peltoniemi 1991; Turner 1992; Pantaleone
1992a; Mukhopadhyaya and Gandhi 1992; Raelt and Sigl 1993).
Moreover, if e mixes with a sterile neutrino species, conversion into
this inert state leads to the loss of energy and lepton number from the
inner core of a SN. The observed SN 1987A neutrino signal may thus
be used to constrain the allowed range of masses and mixing angles
(Kainulainen, Maalampi, and Peltoniemi 1991; Raelt and Sigl 1993;
see also Shi and Sigl 1994).
These applications are discussed in Sects. 9.5 and 9.6 below. A
simple estimate of the rate of avor conversion and the emission rate
of sterile neutrinos from a SN core requires not much beyond the approximate rate 12 sin2 2 . However, a proper kinetic treatment of the
evolution of a neutrino ensemble under the simultaneous action of oscillations and collisions is an interesting theoretical problem in its own
right. Notably, it is far from obvious how to treat degenerate neutrinos in a SN core because the dierent avors will suer dierent Pauli
blocking factors. Does this eect break the coherence between mixed
avors in neutral-current collisions?
The bulk of this chapter is devoted to the derivation and discussion of a general kinetic equation for mixed neutrinos (Dolgov 1981;
Rudzsky 1990; Raelt, Sigl, and Stodolsky 1993; Sigl and Raelt 1993).
This equation provides a sound conceptual and quantitative framework
for dealing with various aspects of coherent and incoherent neutrino
313
9.2
9.2.1
314
Chapter 9
Even coherent or partially coherent density matrices can be diagonalized in some basis. The interactions with the background medium
will also be diagonal in some basis; for neutrinos, this is the weak interaction basis. If the density matrix and the interactions are diagonal in
the same basis, there is no decoherence eect. For example, a density
matrix diagonal in the weak interaction basis implies that there is no
relative phase information between, say, a e and a and so collisions
which aect e s and s separately have no impact on the density
matrix. However, a density matrix which is diagonal in the mass basis,
assumed to be dierent from the weak interaction basis, will suer a
loss of coherence in the same medium.
All told, the loss of coherence is given by a shrinking of the length
of P. More precisely, only the component PT is damped which represents the part transverse to the interaction basis, i.e. which in the
interaction basis represents the o-diagonal elements of . Thus, in the
presence of collisions the evolution of P is given by (Stodolsky 1987)
= V P DPT .
P
(9.1)
The rst 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 eld V is
2
V=
tosc
sin 2
0 ,
cos 2
(9.2)
where is the mixing angle in the medium and tosc the oscillation
period. They are given in terms of the neutrino masses and momentum, 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 eld V, combined with a shrinking of the length of P to
zero. This nal state corresponds to = 21 where both avors are
equally populated, and with vanishing coherence between them. For
1
D = t1
osc and sin 2 = 2 the evolution of the avors 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
315
Matrix of Densities
Which quantity is supposed to replace the previous single-particle density matrix as a means to describe a possibly degenerate neutrino
ensemble? For unmixed neutrinos the relevant observables are timedependent occupation numbers fp for a given mode p of the neutrino eld. They are given as expectation values of number operators
np = ap ap where ap is a destruction operator for a neutrino in mode p
and ap the corresponding creation operator. The expectation value is
with regard to the state | of the entire ensemble. For several avors
it is natural to generalize the fp s to matrices p = (p) of the form49
Strictly speaking (p) is dened by aj (p)ai (p ) = (2)3 (3) (p p )ij (p) and
similar for (p). Therefore, the expectation values in Eq. (9.3) diverge because they
involve an innite factor (2)3 (3) (0) which is related to the innite quantization
volume necessary for continuous momentum variables. In practice, this factor always drops out of nal results so that one may eectively set (2)3 (3) (0) equal to
unity.
49
316
Chapter 9
(Dolgov 1981)
(9.3)
where ai (p) and ai (p) are the destruction and creation operators for
neutrinos of avor i in mode p while b is for antineutrinos which otherwise are referred to by overbarred quantities. The reversed order of the
avor indices in the denition of (p) guarantees that both matrices
transform in the same way under a unitary transformation in avor
space. Also, for brevity | . . . | is always written as . . ..
The diagonal elements of p and p are the usual occupation numbers while the o-diagonal ones represent relative phase information.
In the nondegenerate limit, up to a normalization p plays the role of
the previously dened single-particle density matrix. Therefore, the
p s and p 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 Raelt 1993). It remains to derive an equation of motion
which in the appropriate limits should reduce to the previous precession equation, to a Boltzmann collision equation, and to Stodolskys
damping equation (9.1), respectively.
9.2.3
(t, x) =
(9.4)
More precisely, ap is an annihilation operator for negative-helicity neutrinos of momentum p while bp is a creation operator for positivehelicity antineutrinos. The Dirac spinors up and vp refer to massless negative-helicity particles and positive-helicity antiparticles, respectively; the spinor normalization is taken to be unity. For n avors,
ap and bp are column vectors of components ai (p) and bi (p), respectively. They satisfy the anticommutation relations {ai (p), aj (p )} =
{bi (p), bj (p )} = ij (2)3 (3) (p p ).
In the massless limit and when only left-handed (l.h.) interactions
are present one may ignore the right-handed (r.h.) eld entirely. However, in order to include avor mixing one needs to introduce a n n
317
0p p2 + M 2
)1/2
(9.5)
H0 =
dp
n [
(9.6)
i,j=1
(9.8)
(9.9)
Interactions with a medium are introduced by virtue of a general interaction Hamiltonian Hint (B, ) which is a functional of the neutrino
eld and a set B of background elds; specic cases will be discussed
in Sects. 9.3 and 9.4 below. The equation of motion for p is found from
318
Chapter 9
p (t) = i 0p , p (t) + i
(9.10)
and an analogous equation for p (t). These equations are exact, but
they are not a closed set of dierential equations for the p and p . To
this end one needs to perform a perturbative expansion.
To rst order one may set the interacting elds B(t) and (t) on the
r.h.s. of Eq. (9.10) equal to the free elds50 B0 (t) and 0 (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 Wicks theorem and ignoring
fast-varying terms such as b b it can be reduced to an expression which
contains only p s and p s. The result gives the forward-scattering or
refractive eect 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
(t) = (B(t), (t)) which is a functional of B and is to rst order
(t) = 0 (t) + i
t
0
0
dt Hint
(t t ), 0 (t) ,
(9.11)
0
where 0 and Hint
are functionals of the freely evolving elds B0 (t)
and 0 (t). Applying this general iteration formula to the operator
= [Hint (B, ), p ] which appears on the r.h.s. of Eq. (9.10) one
arrives at
p (t) = i 0p , p (t) + i
t
0
dt
0
Hint
(t), 0p
]
[
0
0
Hint
(t t ), Hint
(t), 0p
]]
, (9.12)
and similar for p (t). The second term on the r.h.s. is the rst-order
refractive part associated with forward scattering. The second-order
term contains both forward- as well as nonforward-scattering eects.
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, (0) = 0 (0) etc. are taken as initial
conditions for the interacting elds. Also, the mass term is ignored in the denition
of 0 ; its eect is included only in the rst 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)eipt etc. for all avors with p = |p|. This implies that
0 (p), which are constructed from the free as and bs, are
the operators 0 (p) and
time independent.
319
Because all operators on the r.h.s. of Eq. (9.12) are free the expectation values in the rst- and second-order term factorize between the
neutrinos and the medium. This leads one to equations for p (t) and
p (t) which on the r.h.s. involve only p (t) and p (t) as well as
0p
and 0p 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. Physically this amounts to the restriction that the neutrino collision rate is
small enough that multiple-scattering eects 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 prescribed, 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 innity while setting p (t) equal to p (0) =
0p .
This leads to
[
p (0) = i 0p , p (0) + i
1
2
0
Hint
(0), 0p
dt
]
[
0
0
Hint
(t), Hint
(0), 0p
]]
, (9.13)
+
and similar for p . Here, 0 dt . . . was replaced by 12
dt . . ..
The dierence between these expressions corresponds to a principlepart 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 interaction model yields the desired set of dierential equations for the p s
and p 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 collision occurs. This assumption corresponds to molecular chaos in the
derivation of the usual Boltzmann equation.
320
9.3
9.3.1
Chapter 9
Neutral-Current Interactions
Hamiltonian
In order to make Eq. (9.13) explicit one must use a specic model for
the interactions between neutrinos and the medium. To this end I begin
with fermions which interact by virtue of an eective neutral-current
(NC) Hamiltonian,
HNC
GF 3
d x Ba (x) (x)Ga (1 5 )(x) .
=
2 a
(9.14)
= 21 e (1 5 )e . With the
r.h. species, a = L or R, so that BL,R
standard-model couplings given in Sect. 6.7.1 one nds in the weak
interaction basis GL = 2 sin2 W + 3 and GR = 2 sin2 W .
For the calculations it is convenient to write Eq. (9.14) in momentum space,
HNC
GF
=
dp dp Ba (p p ) p Ga (1 5 )p ,
2 a
(9.15)
(9.16)
321
Neutrino Refraction
na [Ga , p ] ,
(9.17)
i p = (0p + 2 GF N ), p ,
[
]
(9.18)
i p = (0p 2 GF N ), p ,
where N = diag(ne , 0, 0) in the avor basis (electron density ne ). For
two avors this is equivalent to the previous precession formula. Notably, the and oscillation frequencies are shifted in opposite directions relative to the vacuum energies.
Neutrino-neutrino interactions make an additional contribution to
the refractive energy shifts, i.e. to the rst-order term in Eq. (9.13).
After the relevant contractions one nds51 (Sigl and Raelt 1993)
Sp
2 GF
dq GS (q q )GS + GS Tr (q q )GS
]}
.
(9.19)
322
Chapter 9
neutrinos are important only in young SN cores where one may ignore
antineutrinos. With the total neutrino matrix of densities
dp p ,
(9.20)
Kinetic Terms
The refractive term (rst order) of Eq. (9.13) is just a sum over dierent
medium components, whereas the collision term (second order) in general contains interference terms between dierent target species. However, if they are uncorrelated, corresponding to Ba Bb = Ba Bb
for a = b, these interference terms only contribute to second-order
forward-scattering eects which are neglected. The collision term is
then an incoherent sum over all target species so that in the following
one may suppress the subscript a for simplicity.
323
p,coll =
1
2
(9.23)
(9.24)
where the energy transfer 0 can be both positive and negative. In the
ultrarelativistic limit the neutrino tensor can be written as
N = 12 (U U + U U U U g i U U ),
(9.25)
and p (1 p ) .
(9.26)
324
Chapter 9
dp WP ,P fp (1 fp ) WP,P fp (1 fp )
p =
dp W (P , P ) .
(9.28)
325
(9.29)
where terms nonlinear in p have disappeared even though the neutrinos may still be degenerate.
Eq. (9.29) is more transparent in the case of two-avor mixing where
one may write p = 12 fp (1 + Pp ) and G = 12 (g0 + G ). The
total occupation number fp is conserved while the polarization vector
is damped according to
p,coll = 1 p G (G Pp ).
P
2
(9.30)
(9.31)
Thus one naturally recovers Stodolskys damping term Eq. (9.1) with
D = 12 p |G|2 . For e and and if one writes G = diag(ge , g ) in
the weak interaction basis D = 21 p (ge g )2 . This representation
reects that the damping of neutrino oscillations depends on the difference of the scattering amplitudes: D is the square of the amplitude
dierence, not the dierence of the squares. If one avor does not
scatter at all, D is half the scattering rate of the active avor.
Collisions thus lead to chemical equilibrium as discussed in the introduction to this chapter and as shown in Fig. 9.1. However, the
simple exponential damping represented by Stodolskys 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 avor ones. In a more general case the evolution is more complicated. In particular, collisions usually lead to a
transient avor polarization in an originally unpolarized ensemble if the
momentum degrees of freedom were out of equilibrium. Still, the neutrinos always move toward kinetic and chemical equilibrium under the
action of the collision integral Eq. (9.22) in the sense that the properly
dened free energy never increases (Sigl and Raelt 1993).
9.3.5
Weak-Damping Limit
Even for two-avor mixing the general form of the collision integral
Eq. (9.22) remains rather complicated. However, for the conditions
of a SN core one may apply two approximations which signicantly
simplify the problem. First, one is mostly concerned with the evolution
326
Chapter 9
p = i[p , p ] +
1
2
dp WP P Gp G(1 p )
WP P p G(1 p )G + h.c. .
(9.32)
with vp (sp , 0, cp ).
(9.33)
Here,
sp sin 2p
and cp cos 2p
(9.34)
ep =
fpe
0
0
fpx
1
t
2 p
(fpe
fpx )
0 1
,
1 0
(9.35)
327
1
4
]}
(9.37)
where
w(pa pb ) WP P fpb (1 fpa ) WP P fpa (1 fpb )
(9.38)
(Raelt and Sigl 1993). The equation for fpe is the same if one exchanges
e x everywhere. In the absence of mixing sp = tp = 0, leading to
the usual collision integral for each species separately.
If x is neither nor but rather some hypothetical sterile species
its coupling constant is gx = 0 by denition. In this case the collision
integral simplies to
fpx = 14 s2p ge2
(9.39)
(9.40)
If in addition the mixing angle is so small that the x freely escape one
may
set fpx = 0 on the r.h.s. so that the integral expression becomes
dp WP P fpe .
328
9.3.6
Chapter 9
Small Mixing Angle
n x =
1
4
n x =
1
4
329
9.4
9.4.1
Charged-Current Interactions
Hamiltonian
GF
=
2
d3 x (x)(x) + h.c. ,
(9.43)
where the neutrino eld is, again, a column vector in avor space with
the entries , = e, , in the standard model. Further, is a row of
Dirac operators representing the medium. In the interaction basis
carries the lepton number corresponding to the avor . For example,
in a medium of nucleons and electrons the eld e corresponding to the
electron lepton number can be written for standard-model couplings as
e = (1 5 )e n (CV CA 5 )p ,
(9.44)
where p , n , and e are the proton, neutron, and electron Dirac elds,
respectively, while CV = 1 and CA = 1.26 are the dimensionless CC
vector and axial-vector nucleon coupling constants.
9.4.2
Kinetic Terms
One may now insert HCC into Eq. (9.13) in order to derive the explicit
CC collision integral for the evolution of p and p . The operators
violate the lepton number L corresponding to avor . Therefore,
= 0 at all times if the medium is in an eigenstate of L ( = e, ,
, or additional exotic avors). This assumption implies that the CC
interaction Eq. (9.43) does not contribute to refractive eects given by
the rst-order term in Eq. (9.13).
330
Chapter 9
P
=
A =
1
2
G2F
1
2
G2F
dt ei0 t (, t) (, 0) ,
dt ei0 t Tr (, 0) (, t)
(9.45)
These expressions are dened for both positive and negative energy
transfer 0 because PP
plays the role of an absorption rate for antineutrinos with physical (P0 > 0) four momentum while AP plays
Pe
1
P 0
2
0
P
0
0 .
(9.46)
(9.47)
Weak-Damping Limit
The meaning of Eq. (9.47) becomes more transparent if one makes various approximations which are justied for the conditions of a SN core.
As discussed in Sect. 9.3.5 one may ignore the antineutrino degrees
331
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 nds for two avors (Raelt and Sigl 1993)
fpx = (1 fpx ) PPx fpx AxP
[
(9.49)
(9.50)
332
9.5
9.5.1
Chapter 9
As a rst application of the formalism developed in the previous sections consider two-avor mixing between e and another active neutrino
species x = or (vacuum mixing angle 0 ). In a SN core immediately after collapse electron lepton number is trapped and so the e s
have a large chemical potential on the order of 200 MeV. The trapped
energy and lepton number diuses out of the SN core and is radiated
away within a few seconds. Will x achieve equilibrium with e on this
time scale and thus share the large chemical potential? For the avor this would also imply the production of muons by the subsequent
charged-current absorption of so that the lepton number would be
shared between e, , e , and .
If the mixing angle in the medium were not small, avor conversion
would occur about as fast as it takes to establish equilibrium. In
this case a detailed calculation is not necessary so that one may focus
on the limit of small mixing angles. In addition the oscillations are
fast which allows one to use Eqs. (9.41) and (9.50). Moreover, the
medium properties are assumed to be isotropic so that the production
and absorption rates P e and Ae of e s depend only on their energy E.
Also, the transition rate WP P for the scattering of a neutrino with four
momentum P to one with P may be replaced by an angular average
which depends only on the energies E and E . Altogether one nds a
rate of change for the x density of
n x =
1
4
1
4
a
a
a
2 e
x
dp dp WEE
(gx sp ge sp ) fp (1 fp )
333
GR =
2 sin2 W
0
0
,
2 sin2 W
(9.53)
where sin2 W 41 will be used. Hence the eective NC coupling constants are dierent for e and x interacting with l.h. electrons while
they are the same for r.h. ones.
The e Fermi sea is very degenerate. With regard to the neutrino
distributions one may thus use the approximation T = 0 so that neutrino occupation numbers are 1 below their Fermi surface, and 0 above.
Because the chemical potential e of the e population exceeds x , and
because neutrinos can only down-scatter in the T = 0 limit, the term
proportional to fpx (1 fpe ) vanishes. Moreover, the detailed-balance
requirement (1 fpe )PEe = fpe AeE implies PEe = 0 for E > e where
fpe = 0. Then altogether
n x
(9.56)
(9.57)
where 14 is the density in units of 1014 g cm3 and as usual Yj gives the
abundance of species j relative to baryons. The approximate parameter
334
Chapter 9
|E | = 0
if E /E 1 (large m2 ),
E /E
if E /E 1 (small m2 ).
(9.58)
n x = 02
(9.59)
n x =
9.5.2
0 m2
2 2GF ne
)2
PEe
+
dE
2 2
x
e
dE
a
a
2
a (gx E ge E)
.
WEE
4 4
(9.60)
In order to evaluate these integrals one rst needs the production rate
PEe of e s with energy E due to the CC reaction p + e n + e .
The leptons are taken to be completely degenerate, the nucleons to
be completely nondegenerate. Because they are also nonrelativistic
the absorption of an e produces a e of the same energy. Therefore,
PEe = E np where np is the proton density and E = (CV2 +3CA2 )G2F E 2 /
is the CC scattering cross section for electrons of energy E. Here,
CV = 1 and CA = 1.26 are the usual vector and axial-vector weak
couplings. Altogether one nds
PEe =
CV2 + 3CA2 2 2
GF E np .
(9.61)
335
For l.h. electrons the coupling strengths are dierent, and they are
relativistic so that recoil eects are not small. However, they are degenerate 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 eects. The transition
rates were worked out in detail by Raelt and Sigl (1993) for entirely
degenerate leptons. They found
G2F 2e (E E )E
,
3
E2
G2 2 (E E )E
F e
,
3
E 2
R
WEE
L
WEE
(9.62)
Given enough time the s will reach the same number density as the
e s. Therefore, it is most practical to discuss the approach to chemical
equilibrium in terms of a time scale
d
ln
dt
ne nx
ne
(9.63)
02
3G2F
np 2e (CV2 + 3CA2 ) Fp + e Fe ,
5
e
(9.64)
(9.65)
1 3
) (1
12
)3 /(1 3 ),
(9.66)
336
Chapter 9
and numerically
3G2F
np 2e = 1.0109 s1
5
Yp
14
10 g cm3
)5/3 (
e
e
)2
(9.67)
02
(m2 )2
(CV2 + 3CA2 ) Fp + e Fe ,
8np
e
(9.68)
where Fp = 1 and
Fe =
9 1
137
40
+ 43 (1 + 4 ) log + 53 + 3
8
(1 3 )
3
101 4
120
15 5
(9.69)
1014 g cm3
Yp
)(
(m2 )1/2
1 keV
)4
(9.70)
337
Fig. 9.3. Contour plot for log( 02 ) with in seconds according to Eqs. (9.64)
2 = 4, and / = 1. (Adapted
and (9.68), taking Fe = 0, Fp = 1, CV2 +3CA
e
e
from Raelt and Sigl 1993.)
sin 20 >
2
for m2 <
(100 keV)
2
for m2 >
(100 keV)
(9.71)
338
Chapter 9
9.6
339
Because the x are assumed to escape freely one may set fpx = 0 on
the r.h.s. of these equations. Moreover, for the x 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 nds
(
fpx
PEe
1 2
s
4 p
(gea )2
dp
WEa E
fpe
(9.72)
fpx
PEe
1 2
s
4 p
(gea )2
E 2 WEa E
dE
2 2
dE fpx
and
Q =
(9.73)
dE fpx E
(9.74)
the volume loss rates for lepton number and energy can be easily determined.
2
Turn rst to the case m2 >
(100 keV) so that one may use the
vacuum mixing angle in a SN core. One must now include both the CC
process ep nx as well as the NC process e N N x which is not
suppressed because for a sterile x there is no destructive interference
eect. For nondegenerate nucleons and using CV2 + 3CA2 4 for both
CC and NC processes one easily nds from the results of Sect. 9.5.2
n L = 14 sin2 20
2G2F
nB (Yp + 14 ) 5e ,
5 3
(9.75)
(9.76)
where
1 = 35 (36)1/3 G2F nB = 7.71010 s1 15 .
5/3
5/3
(9.77)
340
Chapter 9
This remark is relevant in the context of recent speculations about the existence
of a 34 MeV sterile neutrino (Barger, Phillips, and Sarkar 1995) as an explanation of
an anomaly observed in the KARMEN experiment (KARMEN Collaboration 1995).
Chapter 10
Solar Neutrinos
The current theoretical and experimental status of the Sun as a neutrino source is reviewed. Particle-physics interpretations of the apparent decit of measured solar neutrinos are discussed, with an emphasis
on an explanation in terms of neutrino oscillations.
10.1
Introduction
(10.1)
which proceeds through a number of dierent reaction chains and cycles (Fig. 10.2). With a total luminosity of L = 3.851033 erg s1 =
2.41039 MeV s1 , the Sun produces about 1.81038 s1 neutrinos, or
at Earth (distance 1.501013 cm) a ux of 6.61010 cm2 s1 . While
this is about a hundred times less than the e ux near a large nuclear power reactor it is still a measurable ux which can be used for
experimentation just like the ux from any man-made source.
The most straightforward application of the solar neutrino ux is
a search for radiative decays by measurements of x- and -rays from
the quiet Sun. Because of the long decay path relative to laboratory
experiments one obtains a limit which is about 9 orders of magnitude
more restrictive (Sect. 12.3.1).
A more exciting application is a search for neutrino oscillations.
In fact, the current measurements of the solar neutrino ux are neither compatible with theoretical predictions nor with each other (solar
neutrino problem); all discrepancies disappear with the assumption of
341
342
Chapter 10
Solar Neutrinos
343
344
Chapter 10
tional sensitivity it is a true neutrino telescope and for the rst time
established that indeed neutrinos are coming from the direction of the
Sun. It is perplexing, however, that the measured ux 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 (Soviet-American Gallium Experiment) and GALLEX began to produce
data in 1990 and 1991, respectively. They are radiochemical experiments using the reaction e + 71 Ga 71 Ge + e which has a threshold
of 233 keV. Therefore, these experiments pick up the dominant pp neutrino ux which can be calculated from solar models with a precision of
a few percent unless something is radically wrong with our understanding of the Sun. Therefore, a substantial decit of measured pp neutrinos
would have been a smoking gun for the occurrence of neutrino oscillations. While the rst few exposures of SAGE seemed to indicate
a low ux, the good statistical signicance of the data that have since
been accumulated by both experiments indicate a ux which is high
enough so that no pp neutrinos are reported missing, but low enough
to conrm the existence of a signicant problem with the high-energy
part of the spectrum (beryllium and boron neutrinos).
With four experiments reporting data, which represent three dierent spectral responses to the solar neutrino ux, the current attention
has largely shifted from a comparison between experiments and theoretical ux 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-dened spectral shape. This sort of analysis currently 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 theoretical solar neutrino ux predictions it should be noted that in recent
years there has been much progress in a quantitative theoretical treatment 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
eld 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
There exist vast amounts of data concerning p-modes (periods between 2 min and 1 h) which can be measured from the Doppler shifts
of spectral lines on the solar surface. The main point is that one can
establish a relationship between the multipole order of the oscillation
pattern and the frequency. Because dierent vibration modes probe the
sound speed at dierent depths one can invert the results to derive an
empirical prole for the sound speed in the solar interior. The agreement with theoretical expectations for the square of the sound speed
is better than about 0.3% except in the inner 0.2 R which are not
probed well by p-modes (Christensen-Dalsgaard, Prott, and Thompson 1993; Dziembowski et al. 1994). Such results make it very dicult
to contemplate the possibility that the Sun is radically dierent from
a standard structure. On the other hand, these results are not precise
enough to reduce the uncertainty of solar neutrino predictions which
arise from the uncertainty of the opacity coecients.
The dierences between the solar neutrino predictions of dierent
authors are minimal when identical input parameters are used. Therefore, the expected error resulting from solar modelling is very small,
except that some key input parameters remain uncertain. The dominant source of uncertainty for the boron neutrino ux is the cross
section for the reaction p + 7 Be 8 B + which appears as a multiplicative factor for the solar ux prediction and thus is unrelated to
solar modelling.
A more astrophysical uncertainty are the opacity coecients. Although they are thought to be well known for the conditions in the
deep solar interior, even a relatively small error translates into a nonnegligible uncertainty of the boron ux prediction because of its steep
temperature dependence. Crudely, a 10% error of the opacity coecients translates into a 1% error of the central solar temperature and
then into a 20% uncertainty of the boron ux. Indeed, a measurement of the solar neutrino ux was originally envisaged as a method to
measure precisely the inner temperature of the Sun.
The Sun as a neutrino source naturally has received much attention
in the literature because of its outstanding potential to nally conrm
the existence of neutrino oscillations in nature. Because this book is
not primarily on solar neutrinos I will limit my discussion to what I
consider the most important features of this unique neutrino source,
and the role it plays for particle physics. A lot of key material can be
found in Bahcalls (1989) book on solar neutrinos, in the more recent
Physics Report on the solar interior by Turck-Chi`eze et al. (1993), and
on the Sun in general in the book by Stix (1989).
346
Chapter 10
Fig. 10.2. Reaction chains PPIPPIII 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 line.
Solar Neutrinos
347
10.2
10.2.1
Individual Sources
Reaction
Q (a )
[MeV]
pp 2 H e+ e
pe p 2 H e
3
He p 4 He e+ e
7
Be e 7 Li e
7
Be e 7 Li e
8
B 8 Be e+ e
13
N 13 C e+ e
15
O 15 N e+ e
17
F 17 O e+ e
Total
0.420
1.442
18.77
0.862
0.384
15
1.199
1.732
1.740
a Maximum
b Bahcall
c
c
c
c
c
c
Flux at Earthb
[cm2 s1 ]
Uncertaintyb
[%]
5.91010
1.4108
1.2103
4.6109
5.2108
6.6106
6.2108
5.5108
6.5106
6.51010
+1
1
+1
2
Factor 6
+6
7
+6
7
+14
17
+17
20
+19
22
+15
19
+1
1
348
Chapter 10
nuclei, the usual weak decay by e+ e 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 rst 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 spontaneous ssion 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 generation. Its importance arises entirely from the high energy of the 8 B
neutrinos which are the ones least dicult to measure.
The neutrino-producing reactions are summarized in Tab. 10.1 with
their maximum energies, and with the resulting neutrino ux at Earth
found in the solar model of Bahcall and Pinsonneault (1995).
Apart from small screening and thermal broadening eects, 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
[
dN/dE = A (Q + me E ) (Q + me E )2 m2e
]1/2
E2 F ,
(10.2)
Source
pp
13
N
15
O
Q [MeV]
A [MeV5 ]
0.420
1.199
1.732
193.9
3.144
0.668
Solar Neutrinos
349
It is much more dicult to obtain the spectrum from 8 B decay because the nal-state 8 Be nucleus is unstable against spontaneous ssion
into two particles. Even though several states of 8 Be contribute to
the transition, it is dominated by the 2.9 MeV excitation and so the
neutrino spectrum can be determined with relatively little ambiguity
by folding Eq. (10.2) with the experimental spectrum (Kopysov and
Kuzmin 1968). A more recent and more detailed analysis was performed by Bahcall and Holstein (1986). An analytic approximation to
their tabulated spectrum is
dN/dE = 8.52106 (15.1 E )2.75 E2 ,
(10.3)
(10.4)
where the quality of the t is equally good as that for the 8 B neutrinos.
350
Chapter 10
Fig. 10.4. Normalized spectra of the thermally and Doppler broadened beryllium neutrino lines where E0 is the corresponding laboratory energy (Bahcall
1994). The line shapes involve an integral over a solar model.
Solar Neutrinos
351
352
Chapter 10
Solar Neutrinos
353
While helium and metal diusion increases the neutrino uxes, 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 ux 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 diusion calculated by dierent
authors since 1988. The temperature predictions are almost uniformly
distributed on the interval 15.4015.72 106 K, i.e. these authors agree
with each other on the value 15.56106 K within 1%.
Thus, in spite of dierences in detail there exists a broad consensus
on what one means with a standard solar model. Therefore, the neutrino uxes of the Bahcall and Pinsonneault (1995) model with element
diusion (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 standard assumptions or incorrect input parameters common to all such
models.
10.2.3
a) Opacities
In order to compare the neutrino ux predictions with the experimental
measurements one needs to develop a sense for the reliability of the
calculations. Naturally, it is impossible to quantify the probability
for the operation of some hitherto unknown physical eect that might
spoil the predictions; an error analysis can only rely on the recognized
uncertainties of standard input physics. Very detailed error analyses
can be found for the standard solar models of Bahcall and Pinsonneault
(1992) and of Turck-Chi`eze and Lopes (1993).
As for solar modelling, the dominating uncertainty arises from the
radiative opacities which largely determine the temperature prole of
the Sun. A reduction of the Rosseland mean opacity in the central
region by 10% reduces the temperature by about 1%.
There is not a one-to-one correspondence between the central solar
temperature Tc and the neutrino uxes because it is not possible to
adjust Tc and leave all else equal. Conversely, one must modify some
input parameters and evolve a self-consistent solar model. Still, if one
allows all input parameters to vary according to a distribution determined by their measured or assumed uncertainties one nds a strong
354
Chapter 10
correlation between Tc and the neutrino uxes which allows one to understand the impact of certain modications of a solar model on the
neutrino uxes via their impact on Tc . Bahcall (1989) found
1.2
T
Tc18
for pp,
for 7 Be,
(10.5)
for 8 B.
However, most recently Tsytovich et al. (1995) have claimed that relativistic
corrections to the electron free-free opacity as well as a number of other hitherto
ignored eects reduce the standard total opacity by as much as 5%.
Solar Neutrinos
355
(10.6)
(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 + reaction
are referenced in Tab. 10.3. In the Sun, the most eective 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
Experiment
(10.8)
Bahcall and Pinsonneault (1992, 1995) used this value, i.e. they used
a 1 uncertainty of 9%. Turck-Chi`eze and Lopes (1993) used (22.4
1.3stat 3.0syst ) eV b, i.e. they adopted an uncertainty of 15%.
Solar Neutrinos
357
10.3
Observations
10.3.1
The longest-running solar neutrino experiment is the Homestake chlorine detector which is based on a reaction proposed by Pontecorvo
(1948) and Alvarez (1949),
e + 37 Cl 37 Ar + e
(10.9)
(10.10)
Because of its low threshold, the gallium experiments can pick up the
solar pp neutrino ux. The absorption cross sections as a function of
358
Chapter 10
37 Cl
and
71 Ga
according to
Because the absorption cross sections are steeply increasing functions of energy while the predicted solar neutrino spectrum (Fig. 10.1)
steeply decreases, the low ux of boron neutrinos yields the dominant
contribution to the expected counting rate for chlorine, and a sizeable
contribution to gallium. This is illustrated in Fig. 10.6 where the predicted counting rates from solar neutrinos, integrated between energy
E and innity, are shown as a function of E . These plots correspond
to the lower panel of Fig. 10.1 if the dierential ux is weighted with
the relevant absorption cross section. It is customary to express the
absorption rate per nucleus in solar neutrino units
1 SNU = 1036 s1 ,
(10.11)
Solar Neutrinos
359
37
71
Cl
Ga
pp
pep
0
11.8
16
215
2.4 10,900
73.2 24,300
Be
13
1.7
61.8
15
6.8
116
360
10.3.2
Chapter 10
Chlorine Detector (Homestake)
Solar Neutrinos
361
Table 10.5. Predicted absorption rate (in SNU) by 37 Cl for dierent solar
source reactions. BP92 give theoretical 3 uncertainties.
Diusion
BP95 He, metals
BP92
He
BP92
TL93
pp
pep
0.0
0.0
0.0
0.0
0.2
0.2
0.2
0.2
1.2
1.2
1.2
1.1
Be
7.2
6.2
5.5
4.6
13
0.1
0.1
0.1
0.1
15
0.4
0.3
0.2
0.2
Total
9.3+1.2
1.4
8.0 3.0
7.2 2.7
6.4 1.4
counters must be subtracted, the counting rate attributed to solar neutrinos in a given run is sometimes found to be formally negative. In
those cases, a zero counting rate is adopted. The distribution of counting rates for 99 runs (18117) is shown in Fig. 10.7. The global average
given in Eq. (10.12) is based on a maximum-likelihood analysis which
includes the background. It would not be correct to average the data
points with individual background subtractions to obtain a global average signal.
362
Chapter 10
The recognized systematic errors are 1.5% for the extraction efciency, 3% for the proportional counter eciency, 3% for the cosmic-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-t average for the solar
neutrino ux measurement, derived from runs 18124 is (Lande 1995)
(2.55 0.17stat 0.18syst ) SNU = (2.55 0.25) SNU
(10.12)
10.3.3
The gallium experiments involve far more complicated chemical extraction procedures for the neutrino-produced 71 Ge which is not a noble
gas. The Soviet-American (now Russian-American) Gallium Experiment (SAGE) used at rst 27 tons, later 55 tons of metallic gallium
while the European GALLEX collaboration uses 100 tons of an aqueous gallium chloride solution, corresponding to 30.3 tons of gallium.
The SAGE experiment is located in the Baksan Neutrino Observatory
in Mount Andyrchi, Caucasus Mountains (Russia) while GALLEX is
located in the Gran Sasso tunnel near Rome (Italy). SAGE has been
taking data since January 1990, GALLEX since May 1991. The current results of SAGE were published by Abdurashitov et al. (1994)
and Gavrin (1995), those of GALLEX by the GALLEX Collaboration
(1994, 1995b) and Kirsten (1995).
The half-life of 71 Ge is 11.43 days so that one needs only about
a three-week exposure, allowing for frequent extractions. SAGE has
already accumulated a total of 21 analyzed runs, GALLEX a total
of 39. In Fig. 10.9 the individual counting rates are shown as well
as the global best-t averages and the distribution of counting rates
in 25 SNU bins. For the SAGE data, formally negative rates after
background subtraction are forced to zero. In both cases, recognized
background signals of 78 SNU were subtracted. The average counting
Solar Neutrinos
363
(10.13)
364
Chapter 10
Table 10.6. Predicted absorption rate (in SNU) by 71 Ga for dierent solar
source reactions. BP92 give theoretical 3 uncertainties.
BP95
BP92
BP92
TL93
Diusion
pp
pep
He, metals
He
69.7
70.8
71.3
71.1
3.0
3.1
3.1
3.0
37.7 16.1
35.8 13.8
32.9 12.3
30.9 10.8
Be
13
3.8
3.0
2.7
2.4
15
6.3
4.9
4.3
3.7
Total
137+8
7
131.5+21
17
127+19
16
122.5 7
Solar Neutrinos
10.3.4
365
A water Cherenkov detector like the one located in the Kamioka metal
mine (Gifu prefecture, Japan) is a large body of water, surrounded
by photomultipliers which register the Cherenkov light emitted by relativistic charged particles. Solar neutrinos are detected by virtue of
their elastic scattering on the electrons bound in the water molecules,
+ e e + .
(10.14)
This process has no signicant threshold, although in practice the detection of the electrons is background-limited to relatively large energies.
The Kamiokande detector began its measurement of solar neutrinos
in January 1987 with an eective analysis threshold of about 9 MeV
which was reduced to about 7 MeV in mid 1988. The much larger
Superkamiokande detector, which is scheduled to begin data-taking in
April 1996, will have a threshold of about 5 MeV.
In contrast with the radiochemical detectors which are eectively
based on the charged-current reaction e + n p + e, the elastic
scattering on electrons is sensitive to all (left-handed) neutrinos and
antineutrinos. The cross section is given by the well-known formula
(e.g. Commins and Bucksbaum 1983)
[
de
G2 me E
me
= F
A + B (1 y)2 C y
,
(10.15)
dy
2
E
where GF is the Fermi constant and the coecients A, B, and C are
tabulated in Tab. 10.7. Further,
E E
Te
2E
y
=
and 0 < y <
,
(10.16)
E
E
2E + me
where E and E are the initial- and nal-state neutrino energies while
Te is the kinetic energy of the nal-state electron.
Table 10.7. Coecients in Eq. (10.15) for elastic neutrino electron scattering.
Flavor A
e
e
,
,
(CV
(CV
(CV
(CV
B
+ CA + 2)2
CA )2
+ CA )2
CA )2
(CV
(CV
(CV
(CV
CV = 12 + sin2 W 0.04,
C
CA )2
+ CA + 2)2
CA )2
+ CA )2
CA = 21 .
366
Chapter 10
For E > 5 MeV, i.e. for energies above the detection threshold
of Kamiokande and Superkamiokande, the total cross section is well
approximated by
E
G2F me E
2.4
(A + 13 B) = 9.51044 cm2
1
2
10 MeV
6.2
1
7.1
for e ,
for e ,
for , ,
for , .
(10.17)
(10.18)
2 (me /E ) cos2
.
(1 + me /E )2 cos2
(10.19)
Solar Neutrinos
367
device and thus a true neutrino telescope. Thus, for the rst time
the Kamiokande detector actually proved that neutrinos are coming
from the direction of the Sun. The rst main publication of these
results was by Hirata et al. (1991) from a data sample of 1040 live detector 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 at so that the spectral shape of the incident
neutrino ux 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 interaction 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 E = 9.5 MeV (Suzuki 1995). Within statistical uctuations the agreement is perfect.
368
Chapter 10
Even though the Sun is thought to produce only e s one may speculate that some of them are converted into e s on their way to Earth
by spin-avor oscillations in magnetic elds (Sect. 8.4) or by matterinduced majoron decays (Sect. 6.8). The e s would cause a signal by
the reaction 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 E = 9.5 MeV
(Suzuki 1995).
be used to constrain the e ux. At dierent energies a 90% CL upper limit on the solar e ux relative to 8 B solar neutrinos is shown
in Fig. 10.14 according to Suzuki (1993). The best relative limit is
at E = 13 MeV where the e ux is less than 5.8% of 8 B solar e s
at 90% CL. If the cause of the isotropic background could be reliably
identied these upper limits could be improved accordingly.
370
Chapter 10
Table 10.8. Predicted ux of 8 B neutrinos at Earth. For BP92, the uncertainty is a theoretical 3 error.
Model
Diusion
BP95
BP92
BP92
TL93
He, metals
He
(10.20)
1990
1991
1987
(1996)
(1996)
(1997)
SAGE
GALLEX
Kamiokande II
Superkamiok.
SNO
BOREXINO
(?)
(?)
(?)
(1996)
(?)
(?)
0.814
Scintillator
x e e x (b )
D2 O Cherenkov
e d p p e
x e e x (b )
x d p n x (b )
H2 O Cherenkov
x e e x (b )
...
0.25
(1.44)
5
2.225
Be
...
...
all
not pp
Thresh. Solar
[MeV]
ux
Radiochemical
0.233
71
71
e Ga Ge e
...
...
Radiochemical
e 37 Cl 37 Ar e
Detection
109 cm2 s1
5.2+0.3
0.4
...
...
6.6+0.9
1.1
137+8
7
...
74 14 SNU
77 10 SNU
2.89 0.41
106 cm2 s1
9.3+1.2
1.4
...
...
4.4 1.1
123 7
...
6.4 1.4
Prediction
BP95 TL93
2.55 0.25
SNU
Measurementc
b = , , or .
c Statistical and systematic 1 errors added in quadrature.
brackets anticipated.
x
e
BP95 = Bahcall and Pinsonneault (1995), with helium and metal diusion.
TL93 = Turck-Chi`eze and Lopes (1993), no diusion.
a In
1967
Homestake
(?)
Operationa
from
until
Experiment
Solar Neutrinos
371
372
10.3.5
Chapter 10
Summary
The main features and results of the solar neutrino experiments discussed 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 aspects 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 uxes when
compared with BP92 (no diusion), while BP95 with the inclusion of
helium and metal diusion 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 diusion would also increase the TL93 uxes.
10.4
Time Variations
10.4.1
Day-Night Eect
Seasonal Variation
Because of the ellipticity of the Earths orbit the distance to the Sun
varies during the year from a minimum of 1.4711013 cm in January
to a maximum of 1.5211013 cm in July, i.e. it varies by 1.67% from
its average during the year. Therefore, the solar neutrino ux varies
by 3.3% from average between January and July. This eect is too
small to be observed by any of the present-day experiments.
This variation can be amplied if neutrinos oscillate. Notably, if
the vacuum oscillation length of the monochromatic 7 Be neutrinos is
Solar Neutrinos
373
374
Chapter 10
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 eect.
Solar Neutrinos
375
The signicance level gives approximately the probability that a random shufing of the data gives an equally good or better (anti)correlation. A signicance
level smaller than 1% is considered highly signicant.
376
Chapter 10
and for the central 14 14 which is the region most signicant for the
neutrino ight path to us. The disk-centered magnetic-eld cycle lags
full-disk indicators by about 1 y. Oakley et al. (1994) found a signicance level of 0.001% for an anticorrelation between the Homestake rate
and the disk-centered magnetic ux. Including the ux from increasingly higher latitudes reduced the signicance 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 between 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 ux
(shaded histogram). A certain degree of broadening is certainly compatible 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 result 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 dierent 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 signicant time variation in the Kamiokande
data between 1987 and 1992 which covers a large fraction of the current solar cycle No. 22. Notably, there is no apparent (anti)correlation
with sunspot number (Suzuki 1993). However, a correlation with diskcentered magnetic indicators may yield a dierent resultaccording 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 conrmed 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
Solar Neutrinos
377
GALLEX, or Homestake. The Homestake argon production rate appears to be strongly anticorrelated with solar disk-centered indicators
of magnetic activity, with no evidence for a long-term variation at
Kamiokande. However, the period covered there (19871992) may be
too short, and the Kamiokande rate has not been correlated with diskcentered indicators. No homogeneous analysis of both Kamiokande and
Homestake data relative to solar activity exists at the present time.
It is very dicult to judge what the apparent Homestake anticorrelation with solar activity means. Is there an unrecognized background
which correlates with solar activity such as cosmic rays? Is there a
long-term drift in some aspect of the experiment which then naturally
correlates with other causally unrelated long-term varying phenomena
such as solar activity? Is it all some statistical uke? Or is it an indication of nonstandard neutrino properties, notably magnetic dipole moments which spin precess in the solar magnetic eld into sterile states?
This latter possibility will be elaborated further in Sect. 10.7.
10.5
10.5.1
Boron Flux
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 ux measurements discussed in
Sect. 10.3 exhibit a decit relative to theoretical predictions. The simplest case to interpret is that of Kamiokande because it is sensitive only
to the boron ux. In this case the most uncertain input parameter is
the cross section for the reaction p7 Be 8 B which enters the ux prediction 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 systematic uncertainties. Therefore, one may turn the argument around
and consider the solar neutrino ux measurement at Kamiokande as
another determination of S17 (0).
According to Eq. (10.20) the measured ux is 2.89 (1 0.14) in
units of 106 cm2 s1 while the prediction is 4.4(10.21)S17 (0)/22.4
for TL93 (Turck-Chi`eze 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
S17 (0)/eV b =
(10.21)
to be compared with the world average 22.4 2.1 (Sect. 10.2.3). While
the discrepancy is severe, one would still be hard-pressed to conclude
with a reasonable degree of certainty that the solar neutrino problem
is not just a nuclear physics problem.
10.5.2
Beryllium Flux
Solar Neutrinos
379
Table 10.10. Measured vs. predicted 7 Be and CNO neutrino uxes (in SNU).
Homestake GALLEX/
(37 Cl)
SAGE (71 Ga)
Measurements:
Total
8
B (inferred from
Kamiokande)
pp + pep
(calculated)
Remainder
2.55 0.25
3.15 0.44
77 10
71
0.20 0.01
74 1
0.8 0.5
4 10
1.24
0.48
1.72
37.7
10.1
47.8
BP95 prediction:a
7
Be
CNO
Total
a Bahcall
An Astrophysical Solution?
Be + e 7 Li + e ,
Be + p 8 B 8 Be + e+ + e .
(10.22)
380
Chapter 10
10.6
Neutrino Oscillations
10.6.1
Solar Neutrinos
381
decits in all detectors may give us the rst indication for neutrino
oscillations and thus for nonvanishing neutrino masses and mixings.
Because of the statistically high signicance 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
ux or its detection methods is ignored, i.e. the apparent anticorrelation with solar activity at Homestake is considered to be a statistical
uctuation. In Sect. 10.7 a possible explanation in terms of neutrino
interactions with the solar magnetic eld is considered.
One may take the opposite point of view that the variability at
Homestake is a real eect, and that it is not related to neutrino magnetic moments because that hypothesis requires fairly extreme values
for the dipole moments and solar magnetic elds (Sect. 10.7). An interpretation of the data in terms of neutrino oscillations then becomes
dicult because one admits from the start that unknown physical effects 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
signicance is reduced.
Still, as no one has put forth a plausible hypothesis for a specic
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 signicant anticorrelation with solar activity.
10.6.2
Vacuum Oscillations
382
Chapter 10
1 AU E
.
osc 10 MeV
(10.23)
Solar Neutrinos
383
384
Chapter 10
The solar neutrino problem has become tightly intertwined with the issue of neutrino oscillations thanks to the work of Mikheyev and Smirnov
(1985) who showed that even for small mixing angles one can achieve
a large rate of avor conversion because of medium-induced resonant
oscillations. This MSW eect was conceptually and quantitatively
discussed in Sect. 8.3.
For a practical application to the solar neutrino problem two main
features of the MSW eect are of great relevance. One is the bathtubshaped suppression function of Fig. 8.10 which replaces a constant
reduction factor (short-wavelength vacuum oscillations) or the wiggly
shape of Fig. 10.17 (long-wavelength vacuum oscillations). It implies
that e s of intermediate energy can be reduced while low- and highenergy ones are left relatively unscathed. This is what appears to be indicated by the observations with the beryllium neutrinos more strongly
suppressed than the pp or boron ones.
Another important feature are the triangle-shaped suppression contours in the sin2 2-m2 -plane for a xed neutrino energy (Fig. 8.9).
Each experiment produces its own triangular band of neutrino parameters where its measured rate is reconciled with the standard ux prediction. As the signal in dierent detectors is dominated by neutrinos
of dierent energies, these triangles are vertically oset relative to one
another and so there remain only a few intersection points where all
experimental results are accounted for. Therefore, as experiments with
three dierent spectral responses are now reporting data the MSW solution to the solar neutrino problem is very well constrained.
Solar Neutrinos
385
(10.24)
(10.25)
Both cases amount essentially to the same neutrino mass-square dierence. Other recent and very detailed analyses are by Fiorentini et al.
(1994), Hata and Langacker (1994), and Gates, Krauss, and White
(1995) who nd similar results. In detail, their condence contours
dier because they used other solar models or a dierent statistical
analysis.
A certain region of parameters (dotted area in Fig. 10.19) can be
excluded by the absence of a day/night dierence in the Kamiokande
counting rates which are expected for the matter-induced back oscillations into e when the neutrino path intersects with part of the Earth
at night. (Even though the detector is deep underground in a mine
the overburden of material during daytime is negligible.) The excluded
range of masses and mixing angles is independent of solar modelling as
it is based only on the relative day/night counting rates.
Except for the Kamiokande day/night exclusion region one should
be careful at taking the condence contours in a plot like Fig. 10.19 too
seriously. The solar model uncertainties are dominated by systematic
eects which cannot be quantied in a statistically objective sense. One
of the main uncertain input parameters is the astrophysical S-factor
for the reaction p 7 Be 8 B . Speculating that all or most of the
missing boron neutrino ux is accounted for by a low S17 has the eect
of reducing the best-t sin2 2 of the small-angle solution by about a
factor of 3 while the large-angle solution disappears, at least in the
analysis of Krastev and Smirnov (1994). The best-t m2 , however,
386
Chapter 10
Fig. 10.19. Allowed range of neutrino masses and mixing angles in the neutrino experiments if the ux decit relative to the Bahcall and Pinsonneault
(1995) solar model is interpreted in terms of neutrino oscillations. The experimental data include all summarized in Sect. 10.3. (Plot adapted from
Hata and Haxton 1995.)
Solar Neutrinos
10.7
387
If the anticorrelation between disk-centered magnetic indicators of solar activity and the event rate at Homestake is taken seriously, the only
plausible explanation put forth to date is that of magnetically induced
neutrino spin or spin-avor transitions that were discussed in Sect. 8.4.
It was rst pointed out by Voloshin and Vysotski (1986) that the varying magnetic-eld strength in the solar convective surface layers could
cause a time-varying depletion of the left-handed solar neutrino ux.
A rened discussion was provided by Voloshin, Vysotski, and Okun
(1986a,b) after whom this mechanism is called the VVO solution to
the solar neutrino problem.
Strong magnetic elds may also exist in the nonconvective interior
of the Sun. In principle, they could also cause magnetic oscillations
and thus reduce the solar neutrino ux (Werntz 1970; Cisneros 1971).
However, the amount of reduction would not be related to the magnetic
activity cycle which is conned to the convective surface layers with an
approximate depth of 0.3R 21010 cm = 200,000 km. The oscillation length is given by Eq. (8.53) so that over this distance a complete
spin reversal is achieved for
BT 31010 B kG,
(10.26)
where B = e/2me is the Bohr magneton and BT is the magneticeld strength perpendicular to the neutrino trajectory. Because the
magnetic eld is mainly toroidal, the condition of transversality is automatically satised.
Magnetic spin oscillations in vacuum do not have any energy dependence and so the solar neutrino ux would be reduced by a common
factor for the entire spectrum. However, a large conversion rate is
only achieved if the spin states are nearly degenerate which is not the
case in media where refractive eects change the dispersion relation of
left-handed neutrinos. In the context of spin-avor oscillations, degeneracy can be achieved by a proper combination of medium refraction
and mass dierences (resonant spin-avor oscillations). In this case
the diagonal elements of the neutrino oscillation Hamiltonian involve
energy-dependent terms of the form (m22 m21 )/2E , causing a strong
energy dependence of the conversion probability. Therefore, the measured signals in the detectors as well as the time variation at Homestake
and the absence of such a variation at Kamiokande can all be explained
by a suitable choice of neutrino parameters and magnetic-eld proles
388
Chapter 10
Solar Neutrinos
10.8
389
Neutrino Decay
See Acker and Pakvasa (1994) for references to earlier discussions of neutrino
decay as a potential solution to the solar neutrino problem.
390
Chapter 10
10.9
Future Experiments
10.9.1
Superkamiokande
If some form of neutrino oscillations are the explanation for the measured solar ux decits 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-avored neutrinos, i.e.
the or appearance rather than the e disappearance. Other clear
signatures would be a deformation of the 8 B spectrum or a diurnal or
seasonal ux variation.
The latter cases can be very well investigated with the Superkamiokande 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 ducial 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 ux, as opposed to about 0.3 events/day at Kamiokande.
Fig. 10.20. Expected signal at night relative to the average daytime signal 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 2 = 0.7
and m2 = 1105 eV2 , the small-angle case for sin2 2 = 0.01 and
m2 = 0.3105 eV2 . Also shown are the existing Kamiokande measurements 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
dierence is m2 = 2105 eV2 , for the small-angle examples (broken lines)
it is m2 = 0.6105 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
to be attributed to a small astrophysical S17 factor. In this case, the explanation for the deciency of beryllium neutrinos could not be resolved
by this detector.
10.9.2
Solar Neutrinos
393
BOREXINO
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, Pakvasa, 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); because 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 scintillator 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 E = 250 keV.
The main challenge at implementing this method is to lower the
radioactive contamination of the scintillator, its vessel, and the surrounding water bath to an unprecedented degree of purity. For example, the allowed mass fraction of 238 U of the scintillator is less than
about 1016 g/g. The feasibility of this method is currently being studied 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, surrounded by 3000 tons of water. Optimistically, data taking with this
facility could commence in 1997.
10.9.4
Summary
Chapter 11
Supernova Neutrinos
The general physical picture of stellar collapse and supernova (SN)
explosions is described with an emphasis on the properties of the observable neutrino burst from such events. The measurements of the
neutrino burst from SN 1987A are reviewed. Its lessons for particle
physics are deferred to Chapter 13 except for the issue of neutrino
masses and mixings. Future possibilities to observe SN neutrinos are
discussed.
11.1
11.1.1
Stellar Collapse
396
Chapter 11
Neutrinos with 10 MeV energies cannot resolve the nucleus, causing it to act
as a single scattering center. Because the neutral-current interaction with protons
is reduced by a factor 1 4 sin2 W (weak mixing angle W ) with sin2 W 0.23
(Appendix B), the elastic scattering cross section of a nucleus scales with the square
of the neutron number.
Supernova Neutrinos
397
Fig. 11.1. Schematic picture of the core collapse of a massive star (M >
Neutrino trapping has the eect 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
( equilibrium), causing a degenerate e 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 eective, 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 prole after collapse is shown in Fig. 11.2
The collapse is intercepted when the inner core reaches nuclear density (0 31014 g cm3 ), a point where the equation of state stiens.
Because the inner core collapse is subsonic, the information about the
central condition spreads throughout, i.e. the collapse of the entire homologous core slows down. However, this information cannot propagate
Supernova Neutrinos
399
beyond the sonic point at the edge of the inner core which now encompasses 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 coresee the thick solid line in Fig. 11.1. Assuming 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 stara SN occurs.
This bounce and shock scenario of SN explosions was rst proposed 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 shocks energy so that in typical calculations 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.
11.1.2
After core bounce and the formation of a shock wave the next dramatic step in the evolution of the core is when the outward propagating
shock breaks through the neutrino sphere, i.e. the shell within which
neutrinos are trapped, most eectively by the coherent scattering on
heavy nuclei. As the passage of the shock wave dissociates these nuclei, it is easier for neutrinos to escape. Moreover, the protons newly
liberated from the iron nuclei allow for quick neutronization by virtue
of e + p n + e , causing a short e burst which is often called the
prompt e burst or deleptonization burst (phase No. 2 in Figs. 11.1
and 11.3). However, the material which is quickly deleptonized encompasses only a few tenths of a solar mass so that most of the leptons
remain trapped in the inner core (Fig. 11.2).
At this stage, the object below the shock has become a protoneutron star. It has a settled inner core within the radius where the
shock wave rst formed and which consists of neutrons, protons, electrons, and neutrinos (lepton fraction YL 0.35). The protoneutron
star also has a bloated outer part which has lost a large fraction of
400
Chapter 11
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 e burst at shock break-out. This outer
part settles within the rst 0.51 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 outward motion and has begun to eject the overburden of matter. Therefore, the protoneutron star after about 0.51 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 avors, and at the same
time deleptonizes by the loss of e s. After 510 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 radiated 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 prole (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 prole during the
rst 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 dicult for neutrinos to escape from this compact
object. Still, the steep gradient of lepton number drives an outward diffusion of neutrinos which move toward regions of lower Fermi momentum and thus, of lower degeneracy energy. Therefore, they downscatter, 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 eciently. 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 rst heats
near the core surface, and then the temperature maximum moves inward 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 e
over e is needed to carry away the lepton number. The total energy
is carried away in almost equal parts by each (anti)neutrino avor.
11.1.3
Supernova Explosions
How do supernovae explode? For some time it was thought that the
outer layers of the star were ejected by the momentum transferred from
the outward neutrino ow which is released after the core collapse (Colgate and White 1966). This scenario had to be abandoned after the
discovery of neutral-current neutrino interactions which trap these particles so that they are released only relatively slowly. With the demise
of the neutrino explosion scenario the earlier suggestion of Colgate and
Johnson (1960) of a hydrodynamic shock wave driving the explosion became the standard, the so-called prompt explosion scenario or direct
mechanism. It continued to malfunction, however, because in numerical calculations the shock tended to stall because of energy dissipation
62
I thank David Seckel for explaining this point to me which is not usually stressed
in the pertinent literature.
402
Chapter 11
Goodman, Dar, and Nussinov (1987) proposed that the pair-annihilation process e+ e might be the dominant mode of energy transfer. However, Cooperstein, van den Horn, and Baron (1987) critizised the neutrino emission parameters
of that study while Janka (1991) found that a proper treatment of the phase space
renders this process less ecient than had been originally thought.
64
This unit is sometimes referred to as 1 foe, for ten to the fifty one ergs.
Supernova Neutrinos
403
Convection is also important at transporting energy within the dilute region between the protoneutron star and the stalling shock wave
(Herant, Benz, and Colgate 1992; Herant et al. 1994; Janka and M
uller
1993a, 1994, 1995a,b; Sato, Shimizu, and Yamada 1993; Burrows,
Fig. 11.4. Unsuccessful (upper panel) and successful (lower panel) SN explosion. 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
dierence between the two cases is the adjusted neutrino luminosity from a
central source; it was chosen as L = 2.101052 erg s1 (upper panel) and
L = 2.201052 erg s1 (lower panel), respectively. In the upper case, which
is just below threshold for a successful explosion, the shock displays an interesting oscillatory behavior. (Curves courtesy of H.-T. Janka, taken from
Janka and M
uller 1993a.)
404
Chapter 11
Fig. 11.5. Entropy contours between neutron star and shock wave in a 2dimensional 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
M
uller (1995b) at t = 377 ms after bounce. (Original of the gure courtesy
of H.-T. Janka.)
Supernova Neutrinos
405
Hayes, and Fryxell 1995). For the rst time 2- and 3-dimensional calculations 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 ow.
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 heating that was found in 1-dimensional calculations (Fig. 11.4) is smoothed
out, but neutrino heating still plays a pivotal role at obtaining a successful and suciently energetic explosion.
In summary, then, the current standard picture of SN explosions is
a modication and synthesis of the Colgate and Johnson (1960) shockdriven 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 suciently 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 2224% helium, the rest hydrogen, and a small
trace of other light elements such as lithium. The present-day distribution 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 different processes at dierent sites. It has long been thought that nuclei with A >
70 were predominantly made by neutron capture, notably the s- and r- (slow and rapid) processes (Burbidge et al. 1957;
Cameron 1957; Clayton 1968; Meyer 1994). The site for the occurrence of the r-process has remained elusive for the past three decades,
although many dierent suggestions have been made. The crux is that
406
Chapter 11
Supernova Neutrinos
407
11.2
11.2.1
Overall Features
M
M
)2 (
10 km
.
R
(11.1)
(11.2)
408
Chapter 11
(11.3)
1012 MeV
for e ,
1417
MeV
for e ,
E =
(11.4)
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
410
Chapter 11
,
dE
1 + eE /T
(11.5)
where T is an eective neutrino temperature and an eective degeneracy parameter. This ansatz allows one to t the overall luminosity by
a global normalization factor as well as the energy moments E and
E2 ; ner details of the spectrum are probably not warranted anyway.
Throughout the emission process, e decreases from about 5 to 3, e
from about 2.5 to 2, and , , , from 2 to 0. This eective degeneracy
parameter is the same for , and , , in contrast with a real chemical
potential which changes sign between particles and antiparticles.
Fig. 11.6. Normalized neutrino spectral distribution according to a MaxwellBoltzmann distribution and a Fermi-Dirac distribution with an eective degeneracy parameter = 2, typical for the Monte Carlo transport calculations of Janka and Hillebrandt (1989a,b). The temperatures are T = 13 E
(Maxwell-Boltzmann) and T = 0.832 13 E (Fermi-Dirac with = 2).
Supernova Neutrinos
411
412
Chapter 11
bounce accretion, and the temperature prole of the core after collapse.
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 eect. 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 nal conguration is not
stable and collapses, presumably to a black hole. This must not occur
too early to avoid conict with the duration of the observed SN 1987A
Supernova Neutrinos
413
signal (Sect. 11.3). However, in some cases studied by Keil and Janka
(1995) with an EOS including hyperons this nal collapse occurs so
late (at 8 s after bounce in one example) that black-hole formation is
dicult 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 dicult
to nd combinations of EOS, core mass, accretion rate, and initial
temperatures which t 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 sti EOS, an initial baryonic
core mass of 1.3 M , and an accretion of 0.2 M within the rst 0.5 s.
The evolution of the eective e luminosity and temperature is shown
in Fig. 11.7. After about 1 s the decay of the temperature is t well by
an exponential et/4 with 10 s while the decay of the luminosity
is poorly t by an exponential; it decays approximately as t1 after 1 s.
For later reference, the time-integrated ux (uence) of this model is
shown in Fig. 11.8.
It must be stressed that the cooling behavior (decrease of the average 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
Fig. 11.8. Expected e uence (time-integrated ux) from SN 1987A, assuming a distance of 50 kpc, and taking Burrows (1988) model 55 neutrino
luminosity shown in Fig. 11.7. The solid line is for an assumed MaxwellBoltzmann energy spectrum for a given average neutrino energy, the dashed
line for a Fermi-Dirac spectrum with a degeneracy parameter = 2 as
in Fig. 11.6.
414
Chapter 11
initial neutrino ux may be characterized by a decreasing radius (decreasing ux), yet increasing temperature. For an overview of various
model calculations see Burrows (1990b). One should be careful not to
take details of the time evolution of the temperature and luminosity of
any specic calculation too seriously.
11.3
11.3.1
Supernova 1987A
Sheltons (1987) sighting of a supernova (SN 1987A) in the Large Magellanic Cloud (LMC), a small satellite galaxy of the Milky Way at a
distance from us of about 50 kpc (165,000 ly), marked the discovery
of the closest visual SN since Keplers of 1604. It was close enough
that several underground detectors which were operational at the time
were able to measure the neutrino ux from the core collapse of the
progenitor star, the blue supergiant Sanduleak 69 202. The observed
neutrinos were registered within a few seconds of 7:35:40 UT (universal
Supernova Neutrinos
415
time) on 23 February 1987 while the rst evidence for optical brightening was found at 10:38 UT on plates taken by McNaught (1987)see
Fig. 11.9.
The main neutrino observations come from the Irvine-MichiganBrookhaven (IMB) and the Kamiokande II water Cherenkov detectors,
facilities originally built to search for proton decay, while a less signicant measurement is from the Baksan Scintillator Telescope (BST).
A likely spurious observation is from the Mont Blanc Liquid Scintillator 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 galactic 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
uctuation.
Koshiba (1992) has given a lively account of the exciting and initially 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
conrmed 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.
11.3.2
Neutrino Observations
416
Chapter 11
The relevant neutrino interaction processes in water are elastic scattering on electrons (Sect. 10.3.4), and the charged-current reactions
e p n e+ and e 16 O 16 F e (Arafune and Fukugita 1987; Haxton
1987). The cross section for the e p reaction is given by
=
(
)
G2F
cos2 C CV2 + 3CA2 pe Ee (1 + )
(11.6)
where GF is the Fermi constant and cos2 C 0.95 refers to the Cabibbo
angle. The charged-current vector and axial-vector weak coupling constants are CV = 1 and CA = 1.26, and incorporates small corrections
from recoil, Coulomb, radiative and weak magnetism corrections (Vogel 1984). Further, pe and Ee refers to the positron momentum and
energy. Ignoring recoil eects, the latter is Ee = E mn + mp
E 1.3 MeV; the threshold is 1.8 MeV because the minimum Ee is me .
A general expression for the e 16 O cross section is much more complicated. A simple approximation, taking only the 2 state of the 16 F
nucleus, is
1.11044 cm2 (E /MeV 13)2
(11.7)
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 E . 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 e 16 O cross section rises very fast; it is
then the dominant detection process for e s. Still, the e p reaction is
the absolutely dominant mode of observing SN neutrinos.
The BST detector is lled with an organic scintillator based on
white spirit Cn H2n+2 with n 9. The dominant detection reaction is
also e p ne+ . In addition, elastic scattering on electrons is possible,
and the process e 12 C 12 N e occurs for Ee >
30 MeV.
The trigger eciencies relevant for the three detectors are shown
as a function of the e energy in Fig. 11.11. Analytic t 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 eect.
The ducial volume of Kamiokande II relevant for the SN 1987A observations was 2,140 tons, for IMB 6,800 tons, and for BST 200 tons. It
corresponds to a target of 1.431032 protons at Kamiokande, 4.61032
at IMB, and 1.881031 at BST.
With the e p cross section of Eq. (11.6) and the eciency curves
of Fig. 11.11 one may compute a prediction for the number of events
418
Chapter 11
Supernova Neutrinos
419
Table 11.1. Neutrino burst at the Kamiokande detector (Hirata et al. 1988).
The time is relative to the rst event at 7:35:35 0:01:00 UT, 23 Feb. 1987.
The energy refers to the detected e , not to the primary neutrino.
Event
Time
[s]
1
2
3
4
5
6a
7
8
9
10
11
12
13a,b
14a,b
15a,b
16a,b
a Usually
attributed to background.
after Loredo and Lamb (1995).
b Quoted
Angle
[degree]
Energy
[MeV]
420
Chapter 11
Table 11.2. Neutrino burst at the IMB detector (Bratton et al. 1988).
The time is relative to the rst 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.
Event Time
[s]
Angle
[degree]
Energy
[MeV]
1
2
3
4
5
6
7
8
80 10
44 15
56 20
65 20
33 15
52 10
42 20
104 20
38 7
37 7
28 6
39 7
36 9
36 6
19 5
22 5
0.000
0.412
0.650
1.141
1.562
2.684
5.010
5.582
Table 11.3. Neutrino burst at the Baksan detector (Alexeyev et al. 1987,
1988). The time is relative to the rst 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.
Event
Time
[s]
Energy
[MeV]
0a
1
2
3
4
5
0.000
5.247
5.682
6.957
12.934
14.346
17.5 3.5
12.0 2.4
18.0 3.6
23.3 4.7
17.0 3.0
20.1 4.0
a Usually
attributed to background.
Supernova Neutrinos
421
the nal-state electron energy distribution is broad so that one can infer
only a lower limit to the 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
422
Chapter 11
Fig. 11.14. SN 1987A neutrinos at Kamiokande and IMB, excluding the ones
which likely are due to background.
Supernova Neutrinos
423
Many authors have studied the distribution of energies and arrival times
of the reported events. Probably the most signicant work is that of
Loredo and Lamb (1989, 1995) who performed a maximum-likelihood
analysis, carefully including the detector backgrounds and trigger efciencies. Loredo and Lamb also gave detailed references to previous
works, and in some cases oered a critique of the statistical methodology employed there. Their more extensive 1995 analysis supersedes
certain aspects of the earlier methodology and results.
Because of the small number of neutrinos observed, a relatively
crude parametrization of the time-varying source is enough. Among a
variety of simple single-component emission parametrizations, Loredo
and Lamb (1989, 1995) found that an exponential cooling model was
preferred. It is characterized by a constant radius of the neutrino
sphere, R, and a time-varying eective temperature
T (t) = T0 et/4 ,
(11.8)
so that 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 eective 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 nds that the average e energy of the timeintegrated spectrum is E e = 2.36 T0 if Fermi-Dirac distributions with
424
Chapter 11
Supernova Neutrinos
425
426
Chapter 11
Zatsepin (1968) was the rst to point out that the e burst expected
from stellar collapse oers a possibility to measure or constrain small
neutrino masses. Because a neutrino with mass m travels slower than
the speed of light its arrival at Earth will be delayed by
(
t = 2.57 s
D
50 kpc
) (
10 MeV
E
)2 (
m
10 eV
)2
(11.9)
Supernova Neutrinos
427
(11.10)
428
Chapter 11
Supernova Neutrinos
429
The Kamiokande events have an obvious correlation between energy and direction. The application of Spearmans rank-ordering test (e.g. Press et al. 1986)
gives a condence level of 0.06% where event No. 6 was excluded as background.
Therefore, the Kamiokande data alone show a fairly signicant angular anomaly.
430
Chapter 11
11.4
Neutrino Oscillations
11.4.1
Overview
Supernova Neutrinos
431
432
Chapter 11
11.4.2
Prompt e Burst
The prompt e burst from a core collapse SN can be detected, in principle, by the forward-peaked signal from the elastic e e ee scattering
in a water Cherenkov detector. In the Kamiokande SN 1987A observations, the rst event could have been caused by the prompt e burst,
but naturally one event contains no statistically signicant information.
It could have been caused by e p ne+ and simply happen to point
in the forward direction. Therefore, the main interest in the prompt e
burst is the possibility that it could be observed from a future galactic
SN by the Superkamiokande or SNO detectors which would yield statistically signicant signatures. A possible oscillation of the e burst
into other avors would reduce the number of forward events because
of the reduced -e scattering cross section of non-e avors (Fig. 11.10).
If the neutrino mass hierarchy is normal where the lightest mass
eigenstate is the dominant e admixture, the medium-induced neutrino
refractive index in the stellar mantle and envelope can cause a mass
inversion and thus level crossing between, say, e and in analogy
to the solar MSW eect. Therefore, one may expect resonant avor
conversion of the prompt e burst in a collapsing star as shown by a
number of authors;67 I follow the analysis of Notzold (1987).
When the shock wave breaks through the neutrino sphere and liberates the prompt e burst, the overlaying part of the progenitor star
has not yet noticed the collapse of its core so that the density prole is
given by that of the progenitor star. The electron density is reasonably
well approximated by a simple power law for which Notzold (1987) used
ne 1034 cm3 r73 ,
(11.11)
Mikheev and Smirnov (1986), Arafune et al. (1987a,b), Lagage et al. (1987),
Minakata et al. (1987), Notzold (1987), Walker and Schramm (1987), Kuo and
Pantaleone (1988), Minakata and Nunokawa (1988), and Rosen (1988).
Supernova Neutrinos
433
ber of e s in the burst is of order 1056 , its duration of order 50 ms. Thus,
while it passes it represents a e density of about 1032 cm3 r72 which
9
exceeds the local electron density for r >
10 cm. However, the phasespace distribution of the neutrinos is locally far from isotropic and so
the energy shift involves a factor 1 cos where is the angle between 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. R/r with R the radius of the neutrino sphere. For a large r one thus nds 1 cos 2 (R/r)2 .
Then, with R 107 cm the eective neutrino density is approximately
1032 cm3 r74 , a value which is always smaller than the electron density
Eq. (11.11). Therefore, in the present context one may ignore neutrinoneutrino 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
prole was used while now Eq. (11.11) pertains. Notzold (1987) found
a conversion probability in excess of 50% if
2
9
m2 sin3 2 >
410 eV E /10 MeV,
(11.12)
2
4
assuming that m2 <
310 eV E /10 MeV so that a resonance can
occur outside of the neutrino sphere. The region in the m2 -sin2 2plane (mixing angle ) with a conversion probability exceeding 50% for
E = 20 MeV is shown as a shaded area in Fig. 11.17, together with
the MSW solutions to the solar neutrino problem. For orientation, the
Kamiokande-allowed range for solar neutrinos is also indicated.
The solar small-angle MSW solution would seem to have a small
impact on the prompt e burst from a collapsing star. Thus, the rst
Kamiokande SN 1987A event may still be interpreted as a prompt e ,
and one may well observe nearly the full e burst from a future SN.
It is interesting that the MSW triangle for the prompt e burst
reaches to relatively large neutrino masses. In the 330 eV regime neutrino masses would play an important cosmological role as dark matter
and for the formation of structure in the universe. Such massive neutrinos likely would mix with e . Unless the mixing angle is very small
the appearance of an unoscillated prompt e burst from a stellar collapse would be in conict with a cosmological role of massive neutrinos
(Arafune et al. 1987b).
434
Chapter 11
Fig. 11.17. MSW triangle for the prompt e burst from a stellar collapse. In
the shaded area the conversion probability exceeds 50% for E = 20 MeV,
assuming the electron density prole 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 e s
Supernova Neutrinos
435
A normal mass hierarchy prevents a level crossing among antineutrinos. Still, if the large-angle or the vacuum solution to the solar neutrino problem obtain, the mixing angle would be so large that
the spectral swapping could still be uncomfortably large. Smirnov,
Spergel, and Bahcall (1994) have considered in detail the probability
for swapping the e with the or spectrum. For vacuum oscillations, relevant for small m2 , the fractional exchange of the spectra
is p = 21 sin2 2 (vacuum mixing angle ) with a maximum of p = 0.5,
i.e. the observed spectrum could be as much as an equal mixture of the
primary ones. In general, medium refractive eects must be included,
although for a large range of m2 one may still take p to be independent of energy. Contours for p in the sin2 2-m2 -plane are shown in
Fig. 11.18. In the shaded area (A) the Earth eect is important so that
the amount of conversion is energy dependent and diers between the
Fig. 11.18. Contours for the swap fraction p between the e and the
or uxes from the protoneutron star cooling phase. The matter eect of
the stellar envelope and of the Earth are included. In (A) the Earth eect 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
Shock Revival
The swap fraction of the e with the more energetic or spectrum discussed in the previous section is always very small unless the
mixing angle is very large because of the assumed normal mass hierarchy which prevents resonant conversions. By the same token a swap
of the e with the or spectrum will be resonant for certain mixing parameters and so it can be almost complete even for very small
mixing angles. If the resonance is located between the neutrino sphere
and the stalling shock wave after bounce, but before the nal explosion, the shock would be helped to rejuvenate because the higher-energy
, s are more ecient at transferring energy once they have converted
into e s (Fuller et al. 1992). Of course, approximately equal luminosities in all avors have been assumed.
A typical density prole for a SN model 0.15 s after bounce is shown
in Fig. 11.19; the step at a radius of about400 km is due to the shock
front. A resonance occurs if m2 /2E
ne . On the right scale
= 2GF1/2
of the plot the quantity mres ( 2GF ne 2E )
is shown for E =
10 MeV and Ye = 0.5. A resonance occurs inside of the shock wave
only for neutrino masses in the cosmologically interesting regime of
order 10 eV and above. Fuller et al. (1992) have performed a detailed
numerical calculation of the additional heating eect for (m2 )1/2 =
40 eV; they found a 60% increase of the energy of the shock wave.
The conversion probability between neutrinos is large if the adiabaticity parameter dened in Eq. (8.39) far exceeds unity. Accord-
Supernova Neutrinos
437
8 E
| ln ne |res ,
m2
(11.13)
(11.14)
R-Process Nucleosynthesis
a) Basic Picture
A partial swap of the e cooling ux with the more energetic or
ux can prevent the synthesis of heavy nuclei by the r-process neutron
438
Chapter 11
capture. As discussed in Sect. 11.1.4, the hot bubble between the settled 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 distribution. 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 reactions e n pe and 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 uxes. The system
is driven to a neutron-rich phase because normally the e s are more
energetic than the e 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 e .
If an exchange e , occurs outside of the neutrino sphere the
subsequent e ux is more energetic than the e ux 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 protonrich state, i.e. to Ye > 21 , to be compared with the standard values of
0.350.46. Therefore, the occurrence of such oscillations would be in
conict with r-process nucleosynthesis in supernovae (Qian et al. 1993).
Fig. 11.20. Density prole 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 2E )1/2 which indicates the resonance value for (m2 )1/2
for E = 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 dierence and mixing angle of e with or where a spectral swap would be ecient enough to help explode supernovae (schematically after Fuller et al. 1992), and where it would prevent r-process nucleosynthesis (schematically after Qian et al. 1993; Qian and Fuller 1994).
440
Chapter 11
1
cos R
1
2
(1 cos ) d cos
(R/r)2
d cos
cos R
(11.15)
)
(
L
L
2GF 1 cos
Le
L e
+
.
4r2
Ee E e E E
(11.16)
The tau-avored neutrino contribution has not been included because
it cancels exactly between and . The same is true for the muavored terms before oscillations have taken place. For a test-neutrino
of momentum p one nds numerically
2p V 420 eV2 (10 km/r)2 1 cos
p
51
10 erg s1
L
L
L e
Le
+
.
Ee E e E E
(11.17)
Supernova Neutrinos
441
Fig. 11.22. Mass splitting between e and (equivalently ) of momentum p = 10 MeV caused by the regular medium (the electrons), and
by dierent neutrino avors according to Eq. (11.17). For the electrons the
density prole of Fig. 11.20 was used with Ye = 0.5; in a real SN core Ye is
much lower near the neutrino sphere so that the thick line would increase
less steeply toward the neutrino sphere than shown here. For the neutrinos,
a luminosity in each degree of freedom of 31051 erg s1 was assumed, the
average energies were taken to be Ee = 11 MeV, E e = 16 MeV, and
E = E = 25 MeV, and the neutrino-sphere radius is 11 km. The
signs of the contributions relative to electrons are: + for e and , for
e and . The contribution of cancels exactly against unless a swap
e has taken place.
442
Chapter 11
linear equations of motion for the neutrino density matrix which causes
an o-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 interactions have a relatively small impact on the parameter space where avor
conversion disturbs the r-process. Within the overall precision of these
arguments and calculations, their nal exclusion plot is nearly identical with the schematic picture shown in Fig. 11.21. Qian and Fullers
ndings are corroborated by a study performed by Sigl (1995a).
c) Summary
In summary, there remains a large range of mixing angles where neutrino oscillations between e and or with a cosmologically interesting mass could help to explode supernovae, and yet not disturb
r-process nucleosynthesis. If one assumes a mass hierarchy with e
dominated by the lightest, by the heaviest mass eigenstate, the cosmologically relevant neutrino would be identied with . It is inter3
<
esting that the relevant range of mixing angles, 3104 <
310 ,
overlaps with the mixing angle among the rst and third family quarks
which is in the range 0.0020.005 (Eq. 7.6). Therefore, a scenario
where a massive plays a cosmologically important role, helps to explode supernovae, and leaves r-process nucleosynthesis unscathed does
not appear to be entirely far-fetched. This scenario leaves the possibility open that the MSW eect solves the solar neutrino problem by
e - oscillations.
11.4.6
A Caveat
Supernova Neutrinos
443
11.5
444
Chapter 11
Another possibility is that the SN explosion itself is not spherically symmetric and thus imparts a kick velocity on the neutron star
(Shklovski 1970). Indeed, as discussed in Sect. 11.1.3 SN explosions
likely involve large-scale convective overturns below and above the neutrino sphere which could lead to an explosion asymmetry of a few
percent, enough to accelerate the compact core to a speed of order
100 km s1 , but not enough to account for the typically observed pulsar
velocities (Janka and M
uller 1994).
An interesting acceleration mechanism was proposed by Harrison
and Tademaru (1975) who considered the rotation of an oblique magnetic dipole which is o-center with regard to the rotating neutron star.
The radiation of electromagnetic power is then asymmetric relative to
the rotation axis and so a substantial accelerating force obtains, enough
to cause velocities of several 100 km s1 . The velocity reached should
not depend on the magnitude of the magnetic dipole moment while its
direction should correlate with the pulsar rotation axis. These predictions do not seem to be borne out by the data sample of Anderson
and Lyne (1983) although the more recent observations may be less
disfavorable to the electromagnetic rocket engine. The correlation
between peculiar velocity and pulsar magnetic moment may now be less
convincing (Harrison, Lyne, and Anderson 1993; Itoh and Hiraki 1994).
Another intriguing mechanism rst proposed by Chuga (1984) relies
on the asymmetric emission of neutrinos (neutrino rocket engine).
Recall that the total amount of binding energy released in neutrinos is
about 31053 erg; because neutrinos are relativistic they carry the same
amount of momentum. If the neutron-star mass is taken to be 1 M ,
and if all neutrinos were emitted in one direction, a recoil velocity of
0.17 c = 5104 km s1 would obtain. Thus an asymmetric emission of
1.5% would be enough to impart a kick velocity of 800 km s1 .
Neutrino emission deviates naturally from spherical symmetry if
large-scale convection obtains in the region of the neutrino sphere.
Janka and M
uller (1994) believe that 500 km s1 is a generous upper
limit on the kick velocity that can be achieved by this method. For a
reliable estimate one needs to know the typical size of the convective
cells as well the duration of the convective phase in the protoneutron
star. To this end one needs to perform a fully 3-dimensional calculation.
No such results are available at the present time.
An asymmetric neutrino emission would also obtain in strong magnetic elds because the opacity is directional for processes involving
initial- or nal-state charged leptons, i.e. URCA processes of the type
e + p n + e or e+ + n p + e . The rates for such processes in the
Supernova Neutrinos
445
11.6
Future Supernovae
The neutrino observations from SN 1987A gave us a wealth of information in the sense that they conrmed the broad picture of neutrino
cooling of the compact object formed after collapse. The data were
much too sparse, however, to distinguish between, say, dierent equations of state or dierent assumptions concerning neutrino transport, or
to detect or signicantly constrain neutrino masses and mixing parameters. Some worry is caused by the apparent anomalies of the SN 1987A
data, notably the angular distribution of the secondary charged particles. No doubt it would be extremely important to observe a SN
neutrino signal with greater statistical signicance, or from a greater
distance. What is the prospect for such an observation?
SN neutrinos can be observed in a number of underground detectors which are operational now or in the near future, or which have only
been proposed. An extensive overview was given by Burrows, Klein,
and Gandhi (1992). Of the experiments which will become operational
within the foreseeable future, the upcoming Superkamiokande water
Cherenkov detector (Sect. 10.9) would yield by far the largest num68
446
Chapter 11
Supernova Neutrinos
447
448
Chapter 11
Chapter 12
Radiative Particle Decays
from Distant Sources
If neutrinos, axions, or other low-mass particles had radiative decay
channels, the decay photons would appear as x- or -ray uxes from
stellar sources where these particles can be produced by nuclear or
plasma processes. This chapter is devoted to limits on such decays,
including decays into charged leptons, that are based on observational
limits on photon or positron uxes from stellar sources, notably the Sun
and supernova 1987A. For comparison, laboratory and cosmological
limits are also reviewed.
12.1
Preliminaries
In this chapter will always denote the partial neutrino decay time into radiation while tot is the total decay time if hypothetical invisible channels are included.
1
Then 1 = B tot
with the branching ratio B .
449
450
Chapter 12
on the process e . The corresponding limit based on a comparison between the measured solar neutrino ux and the measured limit
on x- or -rays from the quiet Sun is 7109 s/eV, almost 9 orders of
magnitude more restrictive. Moreover, this method is fully analogous
to a laboratory experiment as it is based on a measured neutrino ux
and a measured upper limit photon ux. To a lesser degree this remark also applies to the even better SN 1987A constraints which are
applicable to all neutrino avors.
In addition, less directly established particle uxes can be used such
as those from the stars in the galactic bulge or from all hydrogenburning stars or supernovae in the universe. Even more indirectly, one
may study the impact of the radiative decay of the cosmic background
sea of neutrinos or axions which are predicted to exist in the framework
of the big-bang theory of the early universe.
Usually, the bounds on thus obtained are presented as limits on the radiative decay time . Even if one does not aim at an immediate theoretical interpretation, however, the signicance of is limited
because it represents a combination of the nal-state phase-space volume and the matrix element. The latter can be expressed in terms
of an eective transition moment e as in Eq. (7.12) of Sect. 7.2.2.
This moment characterizes the interaction strength independently of
phase-space eects, providing a much more direct link between the experimental results and an underlying theory.
A heavy neutrino h with mh > 2me 1 MeV can decay into
e e+ e . This channel is often included in the notion of radiative decays because relativistic charged leptons cause experimental signatures
similar to rays. If this decay proceeds by virtue of a mixing amplitude
Ueh between h and e the rate is given by Eq. (7.9). Therefore, it is
characterized by |Ueh | in a phase-space independent way.
If e is a mixture of dierent mass eigenstates, any e source such
as a power reactor or the Sun produces all components. If their mass
dierences are small one needs to consider in detail the phenomenon of
neutrino oscillations as in Chapter 8. However, if the oscillation length
is much smaller than the distance between the detector and the source,
the neutrino ux can be considered an incoherent mixture of all mass
2
eigenstates; at a reactor this is the case for m2 >
1eV . The ux of
heavy neutrinos from a e source is then given by
Fh (E ) = |Ueh |2 (E ) Fe (E ) .
(12.1)
The velocity (E ) = (1 m2h /E2 )1/2 enters from the phase space of
nonrelativistic neutrinos in the production process. From e sources
451
one can then derive limits on Ueh for any h , 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 h e e e+ decay.
In the following I will discuss radiative lifetime limits from dierent
sources approximately in the order of available decay paths, from laboratory experiments (a few meters) to the radius of the visible universe
(about 1010 light years).
12.2
Laboratory Experiments
12.2.1
(12.2)
where is the angle between the spin polarization vector and the
photon momentum. For Majorana neutrinos the decay is isotropic,
independently of their polarization, and thus = 0. Similarly for
axion decays, a , 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 1.
The decay photon has an energy = m m /2 in the rest frame of
the parent neutrinosee Eq. (7.12). If it is emitted in a direction
with regard to the laboratory direction of motion, the energy in the
laboratory frame is E = (E + p cos )/m . Hence,
E = 21 m E (1 + cos ),
(12.3)
where = p /E = (1 m2 /E2 )1/2 is the neutrino velocity. For a lefthanded parent neutrino the spin is polarized opposite to its momentum
so that cos = cos and dN /d cos = 21 (1 + cos ). For a very
452
Chapter 12
E
1 + 2
,
m E
0 < E < m E .
(12.4)
The decays of Majorana neutrinos or axions ( = 0) produce a boxshaped spectrum while for = 1 it is triangle shaped (Fig. 12.1).
Fig. 12.1. Photon spectrum from the decay of a relativistic neutrino (energy E ) according to Eq. (12.4).
m dE
E
F (E ) =
d
1 + 2
m E
E /m m
F (E )
. (12.5)
E2
453
454
Chapter 12
Fig. 12.2. Spectrum of reactor e s per ssion of 239 Pu and 235 U (von Feilitzsch 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 Gosgen
reactor (Switzerland). The most recent analysis is, again, based on
data taken with a scintillation counter at Gosgen (Oberauer, von Feilitzsch, and Mossbauer 1987). From a comparison of the reactor on
with the reactor o 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 e radiative decay times of /me > 22 s/eV for = 1, 38 s/eV for = 0,
and 59 s/eV for = +1. It was assumed that is massless so that
m = 1 in Eq. (12.5). With Eq. (7.12) the = 1 constraint translates
into a bound on the eective electromagnetic transition moment of
e < 0.092 B m2
eV ,
(12.6)
not a very restrictive limit even if e saturates its upper mass bound of
about 5 eV.
Even this weak limit would cease to apply if e and became
nearly degenerate. Therefore, Bouchez et al. (1988) performed an experiment where they searched for optical decay photons at the Bugey
reactor (France). They excluded a certain region in the parameter
plane spanned by /me and m . Their greatest sensitivity was approximately at m = 2105 where they found /me >
0.04 s/eV.
7
2
<
With Eq. (7.12) this is e 10 B /meV which, unfortunately, is irrelevant as a constraint. Hence, for nearly degenerate neutrinos there
455
(12.7)
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 h e e+ e will become
kinematically possible and probably dominate. The scintillation counters that were used to search for decay photons near a power reactor
are equally sensitive to electrons and positronsfor many purposes
relativistic charged particles may be treated almost on the same footing as rays. Therefore, the same Gosgen data have been analyzed
to constrain the mixing amplitude Ueh (Oberauer, von Feilitzsch, and
Mossbauer 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 h ux from a e source, and
in Eq. (7.9) to obtain the decay probability. Hence, the expected e+ e
ux is proportional to |Ueh |4 .
456
Chapter 12
(12.8)
For the admixture of other mass eigenstates this result may be translated in a fashion analogous to the discussion of reactor neutrinos.
Beam-stop neutrinos from meson decays may also be used to constrain the e+ e decays of heavy admixtures. Because of the larger
amount of available energy one may probe higher masses for h 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.
It is customary to display |Ueh |2 when constraining the mixing parameters of
heavy, decaying neutrinos while one shows sin2 2eh for light, oscillating neutrinos
as in Chapter 8. For easier comparison I always use the mixing angle. Recall that
|Ueh | = sin eh so that for small mixing angles sin2 2eh = 4|Ueh |2 .
71
457
(12.9)
458
Chapter 12
12.3
12.3.1
Electron Neutrinos
Like a terrestrial power reactor, the Sun is a prolic neutrino source except that it emits e s rather than e s. The expected spectrum as well
as the relevant measurements were discussed in Chapter 10. Suce
it to recall that the solar neutrino ux is now experimentally established without a shred of doubt. There remain signicant 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 ne point in the context of the present discussion because the following results depend mostly on the low-energy
pp ux.
One may proceed exactly as in the previous section in order to translate the solar neutrino spectrum shown in Fig. 10.1 into an expected
ux of x- and -rays from the Sun. In Fig. 12.4 I show this ux at Earth
for /me = 10 s/eV (about the laboratory lifetime limit) and for the
values 1 for the anisotropy parameter . The spectrum as shown is
based on the calculated neutrino ux. The shoulders corresponding to
other than the pp neutrinos likely would have to be reduced somewhat.
The magnitude of the photon ux is enormous because the decay path
d is the entire distance to the Sun of 1.51013 cm = 500 s.
The quiet Sun is a signicant source of soft x-rays from the quasithermal emission of the hot corona at T 4.5106 K. The ux measurements of Chodil et al. (1965) are marked as open diamonds in
Fig. 12.4. The ux of decay photons in Fig. 12.4 would outshine the
solar corona by some 4 orders of magnitude! Moreover, the corona
spectrum falls o sharply at larger energies. In the hard x- and soft
-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 balloonborne detectors own many years ago (Frost et al. 1966; Peterson et al.
1966). These upper limits are still far above the estimated albedo (radiation 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 -ray astronomers and so more recent measurements do not
seem to exist.
In order to respect these measured upper limit photon uxes 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
9
liftime limit for e of /me >
710 s/eV (Cowsik 1977; Raelt 1985).
459
Fig. 12.4. Spectrum of photons from the solar neutrino decay e for
the indicated values of the anisotropy parameter . Measurements of the
x-ray emission of the solar corona (open diamonds) according to Chodil et
al. (1965). Upper limit x- and -ray uxes 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 Raelt 1985.)
(12.10)
460
Chapter 12
Fig. 12.5. Excluded parameters for e from the Sun for nearly degenerate neutrino masses (adapted from Raelt 1985).
461
210
5
|Ueh |2 <
210
3106
(mh = 2 MeV),
(mh = 5 MeV),
(mh = 10 MeV).
(12.11)
Because they used the theoretically expected rather than the experimentally measured solar 8 B neutrino ux I have discounted their original 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 h ux is not diminished by invisible decay channels on its way between Sun and Earth, i.e.
the total h (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 h lifetimes exceeding
about 100 s. Such long-lived MeV-mass neutrinos are in conict with
the big-bang nucleosynthesis constraints shown in Fig. 7.2.
462
12.3.3
Chapter 12
New Particles
Besides neutrinos, stars can also produce other weakly interacting particles 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 nal-state -rays. A short glance at the nuclear reaction chains shown in Fig. 10.2 reveals that a particularly useful case is
p + d 3 He + with E = 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 1.71038 s1 . A certain fraction of the particles produced will be reabsorbed or decay within the Sun. If their probability
for escaping is p the Sun emits r p 1.71038 s1 of the new objects.
If the new particle is a scalar boson a it will have a decay channel
a 2. Because this decay is isotropic in as 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 ux is r p q 2.21010 cm2 s1 MeV1 . The upper limit
photon ux shown in Fig. 12.4 at 5.5 MeV is 0.8103 cm2 s1 MeV1
(Peterson et al. 1966). From there, Raelt and Stodolsky (1982) found
the general upper bound r p q < 41014 . They also calculated r, p,
and q for the specic 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 invisible axions instead (Chapter 14).
12.4
Supernova 1987A
12.4.1
463
(12.12)
m
= F
dLMC
dE
E
1 + 2
E
(E )
,
E2
(12.13)
With uence one means the time-integrated ux. In this book I use the symbol F for a dierential particle ux (cm2 s1 MeV1 ), the symbol F for a uence
(cm2 ), and F = dF/dE for a dierential uence which includes spectral information (cm2 MeV1 ).
464
Chapter 12
E2 eE /T
.
2T3
(12.14)
]
m dLMC [
(1
)
e
+
2
E
()
,
1
2T2
(12.15)
dt ex t /tn . Note
465
range (E,1 , E,2 ) for which one nds by integration of Eq. (12.15)
F,1,2
} 1
m dLMC {
(1 ) e + 2 [ E2 () + E3 ()] ,
= F
2T
2
(12.16)
SMM Observations
466
Chapter 12
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 rst 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.)
4.16.4
1025
25100
0.9
0.4
0.6
6.11
1.48
1.84
b Oberauer
For the 10 s and 223.2 s time intervals one can compute an average
ux limit (cm2 s1 ) 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, conrming their consistency.
467
(12.17)
= F B
dE (1 + 2 E /E ) (E )/E . (12.18)
Therefore, using the same Boltzmann source spectrum the photon spectrum is slightly harder. Going through the same steps as before one
nds the upper limits on B as a function of the assumed T and
75
Somewhat stronger constraints found in the literature were based on a less detailed analysis, notably with regard to the spectral dependence and the dependence
on . Published results are /m > 0.831015 s/eV (von Feilitzsch and Oberauer
1988), 1.71015 (Kolb and Turner 1989), 6.31015 (Chupp, Vestrand, and Reppin
1989), and 2.81015 (Bludman 1992).
468
Chapter 12
5
>
Fig. 12.8. Lower limit on /m for m <
40 eV and tot /m 510 s/eV
(most neutrinos pass the Earth before decaying).
Fig. 12.9. Upper limit on the radiative branching ratio B for m <
40 eV
5
<
and tot /m 510 s/eV (most neutrinos decay between SN 1987A and
Earth).
(12.19)
where s = tot /s. Of course, if tot became so short that the neutrinos
would decay while still within the envelope of the progenitor even this
469
weak limit would not apply. This is the case for lab <
Renv 100 s
(envelope radius Renv of the progenitor star) and so with E 20 MeV
5
one needs to require tot /m >
10 s/eV.
12.4.4
470
Chapter 12
The photons which arrive rst are the ones from decays near the
source. Then the limit dD dLMC leads to d = dLMC dD cos lab and
t = tD (1 cos lab ). With cos lab = ( + cos )/(1 + cos ) where
is the angle of photon emission in the parent frame one nds
t=
tD
,
+ x)
2 (1
(12.20)
where = E /m is the neutrino Lorentz factor and x cos . Therefore, photons detected between t and t + dt result from decays at
tD = 2 (1 + x) t during an interval dtD = 2 (1 + x) dt. (Recall
that t is measured after the rst massless neutrinos arrived while tD
is measured after emission at the source.) The number of parent neutrinos diminishes in time as etD /tot with tot the total decay time.
Therefore, the number of photons traversing a spherical shell of radius
dLMC per unit time is
(1 + x) (1+x) t/tot
N (t) =
e
,
(12.21)
The step function is dened by (z) = 0 for z < 0 and (z) = 1 for z > 0.
471
Fig. 12.11. Arrival time t0 of rst decay photons from a parent neutrino
with velocity , taking the envelope radius of the source to be Renv = 100 s.
The curves are marked with the respective values of x = cos , the direction
of photon emission in the neutrino rest frame.
,
dE dt
+
2
Renv
(12.22)
where = E [(1 )]1 and x = (E / 1)/. The photon ux
at Earth is obtained by multiplication with the neutrino uence F of
Eq. (12.12) and integration over a suitable spectrum (E ) of neutrino
energies.
If the neutrinos are suciently long-lived (the exact meaning of
this is quantied below) the exponential can be ignored. If one also
ignores the absorption eect by the progenitor star (Renv = 0), the
472
Chapter 12
time structure of Eq. (12.22) reduces to (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 rst 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 speaking, 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 (m >
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.
12.4.5
As a rst explicit case for the distribution of photon energies and emission angles I take the two-body decay with a massless daughter neutrino and with the dipole angular distribution of Eq. (12.2) with
x = cos = cos for a left-handed parent. This amounts to
f (, x) = 12 (1 + x) ( 21 m )
(12.23)
2E E
1+
p
where
m
tot ,
2E
tenv
m2
Renv .
2E p
(12.25)
Moreover, the ux vanishes if the chosen value for E does not fall
between 12 (E p ), or equivalently, unless
E > E + m2 /4E ,
(12.26)
473
ignored (tenv = 0) the spectrum is shown in Fig. 12.12 for the anisotropy
parameters = 0, 1. Note the dierence to the triangular shape of
Fig. 12.1 for a stationary source. High-energy photons are now enhanced 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 ux 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 E , according to Eq. (12.24) taking p = E and et/ = 1
(relativistic and long-lived parent), and ignoring absorption eects by the
progenitor (tenv = 0).
In order to compare with the GRS uence limits one needs to integrate the expected ux between t = 0 and t = tGRS = 223.2 s.
The function in Eq. (12.24) is accounted for by using tenv as a lower
limit of integration. Integrating also over the neutrino source spectrum
F (E ) yields
F
[
]
tGRS
2E
2E E
= F
dE
1+
I,
m Emin
p
p
(12.27)
where
I
etenv / etGRS /
.
tGRS /
(12.28)
474
Chapter 12
1 etGRS /
,
tGRS /
(12.29)
1010 eV s.
In the relativistic limit and with I = 1 one can easily integrate
Eq. (12.27) with the Boltzmann spectrum Eq. (12.14) and nds
F = F
]
tGRS [
(1 ) + (1 + ) 2 e ,
m
(12.30)
475
<
Fig. 12.14. Lower limit on m for 200 eV <
m 1 MeV, assuming
10
m tot >
10 eV s.
(12.31)
10
assuming m tot >
10 eV s. Note the dierent dependence on meV
relative to the low-mass result Eq. (12.17).
If m tot violates this condition because it is below 1010 eV s means
that the neutrinos decay so fast that the photon burst eectively ends
before the GRS integration time tGRS is over. Then the photon burst is
again short even though the neutrino mass is large. In this case one
can state a limit on B as in Sect. 12.4.3. If tot is only slightly shorter
8
so that m tot <
410 eV s, all photons arrive within tGRS = 10 s
and one may directly apply Eq. (12.19), originally derived for small
neutrino masses. For the narrow region where the photon burst ends
between 10 and 223.2 s this bound must be discounted by about a
factor of 2 because of the less restrictive uence limits for the larger
integration time.
77
The result here is after Oberauer et al. (1993) which is similar to 61018 eV s
of Bludman (1992) but substantially more restrictive than 0.841018 eV s of Kolb
and Turner (1989). These works all refer to the isotropic case ( = 0).
476
Chapter 12
)
(
tGRS
E eE /T
m2 Renv
= F
dE
, (12.32)
E
m Emin
T3
2p tGRS
Emin = max m 1 +
m Renv
2E tGRS
)2 1/2 (
, E +
m2
. (12.33)
4E
Strictly speaking, the uence of massive neutrinos must be calculated by determining their neutrino sphere which is dierent from the massless case. This
problem was recently tackled by Sigl and Turner (1995) by solving the Boltzmann
collision equation by means of an approximation method known from calculations
of particle freeze-out in the early universe. However, because only masses of up
to 24 MeV are presently considered, a precise treatment of the neutrino spectrum
changes the resulting limits only by a small amount.
477
Summary of Limits
In order to summarize the decay limits I begin in Fig. 12.16 with the relevant regimes of m and tot . Above the upper dotted line the neutrinos
live long enough so that most of them pass the Earth before decaying
while below the lower dotted line they decay within the envelope of
the progenitor star. In the areas 1 and 5 the photon burst is short
<
<
(t <
10 s), in 2 and 4 it is intermediate (10 s t 223.2 s),
and in 3 it is long (223.2 s <
t ). The exact boundaries as well as
the relevant constraints are summarized in Tab. 12.2. In Fig. 12.17 the
limits on e are summarized as a contour plot.
If one restricts possible neutrino decays to the radiative channel one
has tot = which depends only on e and m . Then one may use
directly the upper limits on e given in Tab. 12.2 for the areas 13,
depending on the assumed mass. Put another way, the conditions on
tot are then automatically satised.
These limits certainly apply to e as the e burst from SN 1987A has
been measured. The uxes of the other avors were only theoretically
implied. If they have only standard weak interactions they must have
been emitted approximately with the same eciency as e . Large dipole
moments, however, imply large nonstandard interactions: The same
electromagnetic interaction vertex that allows for radiative decays also
allows for scattering on charged particles by photon exchange! For
MeV energies, for example, the scattering cross section on electrons by
regular weak interactions and that by photon exchange are the same
for e of order 1010 B . Hence in the lower left corner of Fig. 12.17
the neutrinos would interact much more strongly by photon exchange
than by ordinary weak interactions, causing them to emerge from higher
layers of the SN core than normally assumed. Their uence and eective
temperature is then much smaller than standard. Put another way, for
10
e >
10 B the above constraints are not self-consistent (Hatsuda,
Lim, and Yoshimura 1988). However, because large dipole moments can
be constrained by other methods (Sect. 7.5.1) a detailed investigation
of their impact on SN physics is not warranted.
478
Chapter 12
Fig. 12.16. Dierent regimes of neutrino masses and total lifetimes referred
to in the text. t is the duration of the burst of decay photons. The
radiative lifetime limits in the areas 16 are summarized in Tab. 12.2.
The discussion of the previous sections focussed on standard neutrinos which are emitted approximately with the same eciency and
similar energies as e s and e s. It is possible, however, that these neu-
479
Areab
Boundariesc
Radiative lifetime
limitc
Transition momentd
e /B
m < 40
3105 < tot m1
15
m1
> 0.810
1.5108 m2
15
m1
> 0.210
0.8108 m2
m > 71018
1.61010 m1
m < 107
4108 < tot m < 9109
5
105 < tot m1
< 310
B < 1.2109
1.4105 m3/2
tot
m < 107
tot m < 4108
5
105 < tot m1
< 310
B < 31010
0.7105 m3/2
tot
5
tot m1
< 10
B < 0.01
1/2
1/2
b Numbered
trinos have Dirac masses and thus right-handed partners which could
be emitted from the inner core of the SN by helicity-ipping 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
11053 erg is the total energy that was used for a standard plus .
Self-consistency requires fx < 1, of course. In addition, one needs the
average energy Ex of the new objects which allows one to dene 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 uence.
480
12.4.7
Chapter 12
Limit on e e+ e
481
e+ e 2
(12.34)
482
Chapter 12
12.4.8
The bounds on Ue3 from reactors, beam stops, or the Sun were based on
a e ux which partially converts into 3 s which subsequently decay.
Because the SN emits about equal numbers of all ordinary neutrino
avors, this approach is obsolete with regard to 3 . However, one may
still consider hypothetical sterile neutrinos which interact only by virtue
of their mixing with e . 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 h ux 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 eciently than the ordinary ones and so the requirement
that enough energy was left for the observed e s from SN 1987A al10
ready gives one the approximate limit |Ueh |2 <
(Sect. 9.6). If
10
this limit is approximately saturated one expects that about as much
energy is carried away by h as by the ordinary avors. 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 another way, the GRS observations do not dramatically improve on the
cooling argument of Sect. 9.6, although they range in the same general
magnitude.
483
Axions
Supernova Energetics
To derive the various GRS limits one had to assume that the radiative
decays occurred outside of the progenitors envelope and so neutrinos
falling into the shaded area in Fig. 12.16 were not accessible to these
arguments. However, in this case the stellar envelope itself serves as a
detector as discussed by Falk and Schramm (1978) many years ago;
see also Takahara and Sato (1986). Supernova observations in general,
and those of SN 1987A in particular, indicate that of the approximately
31053 erg of released gravitational binding energy only a small fraction
on the order of one percent becomes directly visible in the form of the
optical explosion as well as the kinetic energy of the ejecta. In contrast,
even if only one of the neutrino species decayed radiatively within the
progenitor, about 30% of the binding energy would light up!
If the lifetime were so short that the parent neutrinos would never
get far from the SN core one would not have to worry. Therefore, the
critical range of decay times is between the core dimensions of about
30 km = 104 s and the envelope radius of about 100 s. If the neutrinos
had nonradiative decay modes, and if their total laboratory lifetime fell
2
into this range, one could only conclude that B <
10 .
If they decayed only into radiation, and taking E to be 10 MeV,
the quantity /m cannot lie between about 1011 and 105 s/eV (for
heavy neutrinos includes the e+ e channel). For an eective transition moment e , an interval between about 102 and 105 B m2
eV is
excluded, moderately interesting only for large masses. Then, however,
the cosmological limits strongly suggest the presence of nonradiative
decay channels.
484
Chapter 12
12.5
12.5.1
485
Armed with this result we can return to the question raised in Sect. 7.2.2
if a with a mass exceeding 2me is compatible with the cosmological requirement that such particles and their decay products do not
overclose the universe. It turns out that the SN constraints on
e e+ e presented in this chapter exclude this possibility so that
either respects the cosmological mass limit of a few 10 eV or else it
must have fast invisible decay channels which inevitably require interactions beyond the standard model.
In Fig. 12.19 the available constraints on e+ e are summarized. The
SN 1987A bound from the absence of a prompt burst, the cosmological
requirement, and the above limit from galactic positron annihilation
together exclude the entire range of possible masses and lifetimes. The
margins of overlap are so enormous that each of the arguments has
several orders of magnitude to spare for unaccounted uncertainties.
A heavy standard is also excluded on the basis of arguments involving big-bang nucleosynthesis (BBN). The usual limit on the number of eective neutrino degrees of freedom at nucleosynthesis alone
is enough to reach this conclusion (Sect. 7.1.5). Moreover, charged
leptons and secondary photons from the e+ e decay channel would destroy some of the synthesized nuclei (Lindley 1979, 1985; Krauss 1984;
Kawasaki, Terasawa, and Sato 1986). The main virtue of the SN limits
is, therefore, that no reference to BBN is required to exclude a heavy .
486
Chapter 12
Fig. 12.19. Excluded areas of the mass and lifetime if the standard-model
decay e e+ e is the only available channel. The laboratory results
refer to the bounds on sin2 2e3 of Fig. 12.3, translated into a limit on e+ 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.
12.6
(12.35)
487
MeV
E
)2
(12.36)
s
N
(12.37)
which does not depend on the assumed value for E because Eq. (12.35)
and (12.36) both scale with E 2 . This limit applies if the total neutrino
lifetime tot exceeds tU ; otherwise only a limit on the branching ratio
B can be found.
The most prolic stellar neutrino source in the universe are hydrogen-burning stars which produce two e s with MeV energies for
every synthesized 4 He nucleus. Because most of the binding energy
that can be liberated by nuclear fusion is set free when single nucleons
are combined to form 4 He, most of the energy emitted by stars can
be attributed to hydrogen burning. Therefore, it is easy to translate
the optical luminosity density of the universe into an average rate of
neutrino production.
The average luminosity density of the universe in the blue (B) spectral band is about h 2.4108 L,B Mpc3 where L,B is the solar B luminosity. The Sun produces about 11038 e /s and so one arrives at
N e h 2.41046 Mpc3 s1 = 0.81027 cm3 s1 . (Of course, this estimate is relatively crude in that the neutrino luminosity scales directly
with the average bolometric luminosity of a stellar population, but not
precisely with LB .) With Eq. (12.37) this leads to a constraint for e
2
12
7
<
of /me >
510 s/eV, or e 210 B meV .
The (core-collapse) supernovae in the universe are also very prominent neutrino sources, and, more importantly, they are thought to produce MeV neutrinos of all avors. Such SNe do not occur in elliptical galaxies, and their present-day rate in spirals depends sensitively
488
Chapter 12
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 supernova unit is dened by 1 SNu 1 SN per century per 1010 L,B .
Adopting 1 h2 SNu as a representative value and about 51057 neutrinos plus antineutrinos of a given avor per SN yields for each avor N h3 1.31027 cm3 s1 , a rate almost identical to that from
hydrogen-burning stars.79 The radiative lifetime limit is then also identical, except that it applies to neutrinos of all avors.
All of these bounds are weaker than those from SN 1987A. Therefore, decaying stellar neutrinos cannot actually contribute to the observed x- and -ray background.
12.7
Cosmological Bounds
12.7.1
Neutrinos
Multiplying this rate with the age of the universe of about 31017 s and the
speed of light of 31010 cm/s, and using h = 0.5 one nds an estimated present-day
ux at Earth of about 1 cm2 s1 . In a recent detailed study, Totani and Sato
(1995) nd a ux which is larger than this crude estimate by as much as a factor
of 30. The Kamiokande II detector has set an upper limit on the cosmic background
ux of e of about 103 cm2 s1 for eective temperatures in the 34 MeV range
(Zhang et al. 1988). It is conceivable that this background will be measured by
the Superkamiokande detector. Note that at the Kamiokande site the e ux from
power reactors is roughly 1000 times larger than the background ux, except that
it falls o sharply beyond about 10 MeV.
489
(12.38)
490
Chapter 12
491
still play a signicant dynamical role. Recently, such mixed dark matter
scenarios have received much attention where m = 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
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 matter is actually required to solve certain problems, notably the ionization 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 specic: An energy of decay photons of E =
(14.4 0.5) eV and thus a neutrino mass of m = (28.9 1.1) eV
with a radiative lifetime of = (2 1) 1023 s which translates into
e = (6.3 2) 1015 B . 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
diuse extragalactic background light are more reliable in this regard.
While they marginally exclude Sciamas 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
denitely would have to show up if neutrinos were the bulk of the
galactic dark matter (e.g. Sciama 1993b).
12.7.2
Axions
In the early universe, axions are also produced by the relaxation of the coherent
initial eld conguration 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 < 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 a h2 (105 eV/ma )1.175 see Eq. (14.5). While the overall coecient of
this expression is very uncertain it is clear that a = 1 saturates for ma somewhere
between 1 eV and 1 meV. In this range axions never achieved thermal equilibrium.
Chapter 13
What Have We Learned from
SN 1987A?
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
eects could have caused a time delay between photons and neutrinos
or between e s of dierent energy. Third, nonstandard cooling agents
could have shortened the neutrino burst below its observed duration.
These arguments are applied to a variety of specic cases.
13.1
Introduction
In Chapter 11 the neutrino observations from SN 1987A were discussed and it was shown that they agree well with standard theoretical 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 Baksan event clusters observed at 7:35 UT on 27 February 1987 represent
the e component of the neutrino burst from the core collapse of the
SN 1987A progenitor star rather than some other particle ux, or some
other reaction than the expected dominant e p ne+ process.
Accepting this, there is a host of consequences concerning a variety
of fundamental physics issues. The rst and simplest set of arguments
is based on the fact that the e pulse and perhaps the prompt e burst
493
494
Chapter 13
were observed, constraining various mechanisms that could have removed neutrinos from the beam such as decays. Equally important, the
nonobservation of a -ray burst in coincidence with the neutrino burst
constrains radiative decays of neutrinos and other particles (Sect. 12.4).
More intricate arguments involve signal dispersion, either between
photons and neutrinos, between e s and e s, or the intrinsic dispersion
of the e burst, constraining various eects that could cause signal
dispersion such as a nonzero neutrino mass or charge.
Most importantly, the inferred cooling time scale of a few seconds
of the newborn neutron star precludes an ecient operation of a nonstandard cooling agent and thus yields constraints on the emission of
new particles from the SN core, notably of right-handed (r.h.) neutrinos or axions. This line of reasoning is analogous to the energy-loss
argument which for normal stars has been advanced in Chapter 2.
13.2
13.2.1
Fluence
The neutrinos from a SN are expected to consist of two major components: the prompt e burst, and quasi-thermal emission of about equal
total amounts of energy in (anti)neutrinos of all avors. The water
Cherenkov detectors would register the e burst by virtue of the reaction e + e e + e where the scattered electron is strongly forward
peaked, while the cooling signal is registered by e + p n + e+ with
an essentially isotropic e+ signal. Even though both detectors observed
a forward peaked overall signal, it cannot be associated with e -e collisions (Sect. 11.3.5). Most or all of the events are interpreted as e s.
The observation of a e uence (time-integrated ux) roughly in
agreement with what is expected from a stellar collapse precludes that
these particles have decayed on their way from the SN to us, yielding
a constraint on their lifetime of (Frieman, Haber, and Freese 1988)
5
e /me >
610 s/eV.
(13.1)
However, this simple result must be interpreted with care because massive neutrinos are expected to mix. The heavy e admixtures could
decay and may violate this bound.
The e s were not removed by excessive scattering on cosmic background neutrinos, majorons, dark-matter particles etc., leading to constraints on secret interactions (Kolb and Turner 1987). Take the
495
scattering on cosmic background neutrinos as an example. The presentday density of primordial neutrinos is about 100 cm3 in each neutrino and antineutrino avor. With a distance to the Large Magellanic
Cloud of about 50 kpc = 1.51023 cm one has a column density between SN 1987A and Earth of about 1025 cm2 so that the e - cross
section must be less than about 1025 cm2 . If the cosmic background
neutrinos are massless they have a temperature of about 1.8 K and
so E 3T 5104 eV. Because the measured SN neutrinos
have
a characteristic energy of 30 MeV the center of mass energy is
s 200 eV.
The cross-section bound is not particularly impressive compared
with a standard weak cross section of order G2F s 1051 cm2 . However, cross sections have never been directly measured and so the
SN 1987A limit provides nontrivial information. As an example, neutrinos could scatter by majoron exchange, or they could scatter directly on a background of primordial majorons. Kolb and Turner then
found a certain constraint on the neutrino-majoron Yukawa coupling
(Sect. 15.7.2). As another example, the proposition that the solar neutrino ux could be substantially depleted by scatterings on cosmic background particles (Slad 1983) is excluded.
Other particles besides neutrinos may have been emitted from the
SN and could have caused detectable events. Engel, Seckel, and Hayes
(1990) have discussed the case of axions; they can be absorbed in water
by oxygen nuclei, a16 O 16 O , which subsequently produce rays by
decays of the sort 16 O 16 O , 16 O 15 O n , and 16 O 15 N p .
The rays would cause electromagnetic cascades and so they are detectable about as eciently as e . The axion emission was estimated by
identifying their unit optical depth for a given interaction strength in a
simplied model of the SN temperature and density prole. More than
10 extra events would be expected at Kamiokande for an axion-nucleon
Yukawa coupling in the range
3
<
1106 <
gaN 110
(13.2)
which is thus excluded. In the middle of this interval, up to 300 additional events would have been expected. However, axions with couplings in this interval are also excluded by other methods (Sect. 14.4).
13.2.2
Energy Distribution
496
Chapter 13
Prompt e Burst
In the rst publication of the Kamiokande group (Hirata et al. 1987) the second
event was also reported forward; its most probable direction was later revised.
497
No rays in conjunction with the SN 1987A neutrino burst were observed by the solar maximum mission (SMM) satellite which was operational at the relevant time. Therefore, one can derive some of the most
restrictive limits on neutrino radiative decays as detailed in Sect. 12.4.
13.3
Dispersion Eects
13.3.1
498
Chapter 13
so that the speed of light and that of neutrinos are equal to within
(Longo 1987; Stodolsky 1988)
c c
<
2109 ,
c
(13.3)
V [r(t)] dt,
(13.4)
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
< 0.74103 ,
t
(13.5)
where the uncertainty reects the uncertain modelling of the gravitational potential between Earth and SN 1987A.83 This result has been
used to constrain the parameters of a specic model of C- and Pviolating gravitational forces (Almeida, Matsas, and Natale 1989), and
to constrain the parameters of a class of nonmetric theories of gravity
(Coley and Tremaine 1988).
13.3.2
For a recent laboratory experiment which addresses the Lorentz limiting velocity, see Greene et al. (1991), and references there to earlier works. See also the book
by Will (1993).
83
See Will (1993) for a review of many other empirical tests of general relativity.
499
a) Neutrino Mass
So far the transit time of dierent particle species was compared under the assumption of a xed velocity each. However, the most likely
eect of signal propagation over large distances is dispersion due to an
energy-dependent speed of propagation. The most widely discussed84
case is that of a nonzero neutrino mass (Zatsepin 1968). The main
problem at extracting information about the signal dispersion is the
unknown behavior of the source which must be modelled according
to some theoretical assumptions. A particularly detailed discussion is
that of Loredo and Lamb (1989) who found a mass limit of me < 23 eV
(Sect. 11.3.4). In a similar analysis which included the 13% dead-time
eect at IMB, Kernan and Krauss (1995) found me < 20 eV.
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 e would bend the neutrino path in the galactic magnetic eld,
leading to a time delay of
t
e2 (BT dB )2
=
,
t
6E2
(13.6)
where BT is the transverse magnetic eld and dB the path length within
the eld. This leads to a constraint of
(
e <
1 G
31017
e
BT
)(
1 kpc
dB
(13.7)
(Barbiellini and Cocconi 1987; Bahcall 1989). Note that a typical eld
strength for the ordered magnetic eld in the galactic spiral arms is
23 G 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 me from the SN 1987A data were derived, among others, by Abbott,
de R
ujula, and Walker (1988), Adams (1988), Arnett and Rosner (1987), Bahcall 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 fth-force
charge. If electrons, protons, or dark-matter particles also carry such
a charge the bending of the neutrino trajectory in the fth-force eld of
the galaxy would lead to an energy-dependent time delay. This and related arguments were advanced by a number of authors (Pakvasa, Simmons, and Weiler 1989; Grifols, Masso, 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 mediated 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 neutrino background would be essentially a neutral plasma with regard to
the new interaction. The resulting screening eects then invalidate the
SN 1987A argument (Dolgov and Raelt 1995).
Screening eects 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 energymomentum tensor and thus is identical with gravity. The force mediated 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 eect would be weakened for relativistic neutrinos.
In summary, the SN 1987A signal does not seem to carry any simple
information concerning putative nongravitational long-range forces.
d) Fundamental Length Scale
Fujiwara has proposed a quantum eld theory where the velocity of particles increases with energy, leading to an energy-dependent advance of
the arrival times by t/t = 12 (0 E )2 . Here, 0 is a fundamental length
18
scale. Whatever the merits of this theory, a value 0 <
10 cm would
not be in conict with the SN 1987A neutrino signal (Fujiwara 1989).
e) Lorentz Addition of Velocities
If relativistic particles (photons, massless neutrinos) are emitted by a
moving source (velocity vS ) their velocity c in the laboratory frame
should be equal to c (velocity in the frame of the source). The Galilean
addition of velocities, on the other hand, would give c = c + vS . In
general one may assume that velocities add according to c = c + KvS
501
K <
10 derived from the time of ight of decay photons 2
13.4
13.4.1
General Argument
The most intricate way to use SN 1987A as a laboratory arises from the
observed duration of neutrino cooling. While the neutrino luminosity
during the rst few 100 ms until the shock has been revived is largely
powered by accretion and by the contraction and settling of the bloated
outer core, the long tail is associated with cooling, i.e. emission from
the neutrino sphere which is powered by energy originally stored deep
in the inner core. If a direct cooling channel existed for that region,
such as the emission of r.h. neutrinos or axions, the late cooling phase
would be deprived of energy. Put another way, a novel cooling channel
from the inner core would leave the schematic neutrino light curves
of Fig. 11.3 more or less unchanged before about 1 s while the long
Kelvin-Helmholtz cooling phase would be curtailed.
502
Chapter 13
503
504
Chapter 13
In order to estimate the impact of a novel cooling channel on the neutrino signal it is obviously useful to evolve a protoneutron star numerically with the new physics included, and to calculate the expected
neutrino signal for a varying strength of the new eect. Considering
the many uncertainties involved in this procedure one may well ask if
it is not just as reliable to perform a simple analytic estimate.
At about 1 s after core bounce the neutrino luminosity in all six
(anti)neutrino degrees of freedom together is about 31052 erg s1 . The
mass of the object is around 1.5 M = 31033 g so that its average
energy-loss rate is L /M 11019 erg g1 s1 . A novel cooling agent
would have to compete with this energy-loss rate in order to aect the
total cooling time scale signicantly. Therefore, the observed signal
duration indicates that a novel energy-loss rate is bounded by
19
1 1
x <
10 erg g s .
(13.8)
505
Fig. 13.2. Prole of various parameters for the protoneutron
star model S2BH 0 of Keil,
Janka, and Raelt (1995), 1 s
after core bounce. The degeneracy parameters were approximated by N = (EF mN )/T
with EF2 = p2F + m2N and with
the eective nucleon mass.
506
Chapter 13
The prole of various parameters as a function of the mass coordinate is shown in Fig. 13.2 for model S2BH 0 of the cooling calculations of Keil, Janka, and Raelt (1995); it illustrates typical physical
conditions encountered in the core of a protoneutron star during the
Kelvin-Helmholtz phase. For this model, the average value of (/0 )n
with the nuclear density 0 = 31014 g cm3 and of (T /30 MeV)n is
shown in Fig. 13.3 as a function of n.
Fig. 13.3. Average values for (/0 )n with the nuclear density 0 =
31014 g cm3 and of (T /30 MeV)n for the protoneutron star model of
Fig. 13.2.
the degenerate emission rate Eq. (4.10) yields an almost identical result.
Therefore, a simple criterion like Eq. (13.1) is not a bad rst estimate
for the import of a novel energy-loss rate.
13.4.3
Trapping Limit
When the new particles (for example, axions) interact strongly enough,
they will be emitted from a spherical shell where their optical depth is
about unity rather than by volume emission. Again, one is concerned
507
mostly with a time later than 0.51 s where the outer core has settled and the shock has begun to escape. The density of the protoneutron 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
4r2 Tx4 with the Stefan-Boltzmann constant which is g 2 /120 in
natural units with g the eective number of degrees of freedom (2 for
photons). Therefore, one must demand that
1/4
Tx <
,
8 MeV g
(13.9)
(13.10)
2
Because n is a relatively large number such as 37 the criterion a <
3
6
<
yields R >
n (TR /Ta ) . With the requirement Ta 8 MeV and with
>
TR 20 MeV as taken by Turner one nds ga 2107 , not in bad
agreement with what one would conclude from the numerical results
shown in Fig. 13.1.
Still, this argument is rather sensitive to the detailed model assumptions concerning the protoneutron star structure. Also, as axions
contribute to the transfer of energy within the star, a self-consistent
model must take this eect into account. Moreover, for novel fermions
such as r.h. neutrinos one must distinguish carefully between their neutrino sphere (from where they can escape almost freely) and the deeper
508
Chapter 13
region where their energy ux 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 complicated 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 contrary, 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 dimension 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
509
recent numerical study by Keil (1994) who used the same axion emission rates as Burrows, Turner, and Brinkmann (1988) conrmed their
results.
Here, I present the numerical studies of Burrows and his collaborators where axions were assumed to couple with equal strength to protons and neutrons. The axial-vector coupling to nucleons is written in
the form (C/2fa ) 5 a with a model-dependent numerical factor
C, the Peccei-Quinn energy scale fa , the nucleon Dirac eld , and the
axion eld a. Under certain assumptions detailed in Sect. 14.2.3 it can
be written in the pseudoscalar form i ga 5 where ga = CmN /fa
is a dimensionless Yukawa coupling (nucleon mass mN ); Burrows et al.
used C = 21 . All results will be discussed in terms of ga and as such they
apply to any pseudoscalar particle which couples to nucleons accordingly. 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 implemented 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 cooling from an axion sphere was implemented by means of an eective
radiative opacity as discussed in Sect. 4.4. The protoneutron star models are those of Burrows and Lattimer (1986) and of Burrows (1988b).
In the latter study, cooling sequences were presented for dierent equations of state (EOS), and dierent assumptions concerning the mass
and early accretion rate of the stars. A ducial case in these studies is
model 55 with a sti 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 dened 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. Recall from Sect. 11.3.2 that the observed SN 1987A numbers of events
are NIMB = 8 and NKII = 1012, 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 (sti or soft) is indicated. The models were calculated
until 20 s after collapse.
511
512
Chapter 13
rate 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 prole of this /T
is shown in Fig. 13.5. In Fig. 4.8 the axion emission rate was shown as a
function of , 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 uctuation rate may be smaller than the OPE
calculated value, taking one perhaps somewhat to the left of the maximum. Therefore, the true axion emission rate corresponds to the naive
one (dashed line in Fig. 4.8) at /T 35 which at temperatures
around 30 MeV corresponds to around 20% nuclear density.
Fig. 13.5. Prole for the nondegenerate spin-uctuation rate of Eq. (4.7)
in the protoneutron star model S2BH 0 of Keil, Janka, and Raelt (1995)
shown in Fig. 13.2.
(13.11)
13.6
513
One may ask how the neutrino signal from a SN would be modied
if there existed additional light sequential neutrino avors beyond e ,
, and . Of course, the Z decay width measured at CERN already
reveals that there are exactly three sequential neutrino avors (Particle
Data Group 1994); the same conclusion is reached from studies of big
bang nucleosynthesis (e.g. Kolb and Turner 1990).
Burrows, Ressell, and Turner (1990) calculated several protoneutron star cooling sequences, varying the number of avors from 3, the
standard value, to 11. This increases the eciency of energy transfer
within the SN core and also allows for a more ecient radiation from
the neutrino sphere as there are more degrees of freedom. Thus one
expects a shortened signal in the Kamiokande II and IMB detectors,
as well as a reduced number of events because the available energy is
shared between more neutrino degrees of freedom of which mostly the
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 avors (Burrows, Ressell, and Turner
1990). The signal duration is dened as the time it takes to accrue 90% of
the total expected number of events.
514
Chapter 13
e s are detected. These expectations are borne out by the numerical results shown in Fig. 13.6. A doubling of the number of avors is
probably excluded by the observed signal duration.
13.7
Neutrino Opacity
)2
13
310 g cm3
)2
10 MeV
.
T
(13.13)
515
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 dened 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 Raelt 1995.)
refer to a suppression of the neutral-current reactions alone. The modication of the results between those cases is relatively minor, indicating
that the neutral-current interactions represent the dominant opacity
source for the overall cooling time scale.
The increase of the counting rates at the two detectors with decreasing opacities is explained by the neutrino sphere moving to deeper and
hotter layers, yielding larger neutrino energies. The detectors register
mostly e s so that the number of events is relatively sensitive to the
charged-current opacity which aects only the electron avor. NIMB is
particularly sensitive to the e spectrum because of its high threshold.
516
Chapter 13
By the same token, NIMB is quite sensitive to spectral pinching, an effect not included in these calculations where an equilibrium neutrino
transport scheme was used. Therefore, the total number of events,
notably at IMB, is a poor measure to characterize the neutrino signal.
In the calculations of Keil, Janka, and Raelt (1995) the mass of
the initial neutron-star model as well as its temperature prole and the
equation of state were varied. While such modications cause changes
in the predicted signal durations and event counts, none of these parameters appears likely to be able to compensate for an extreme suppression of the neutrino opacities. The SN 1987A neutrino signal excludes
a suppression eect stronger than, say, a >
0.5.
These ndings are in agreement with those of the previous section
where the number of neutrino avors had been increased. Essentially
that procedure amounted to increasing the eciency of neutral-current
energy transfer and so it is not very dierent from the decreased opacities used here.
Therefore, it appears that the standard opacities which ignore
fast spin uctuations provide a reasonable representation of what is
observed. This result appears to imply that the spin-uctuation rate
does not exceed O(T ) in a SN coresee Sect. 4.6.7 for a discussion
of these matters. However, it is surprising that the best t is achieved
by the naive opacities because the spin-uctuation eect is only one
reason to expect reduced axial-vector opacities. Other reasons include
reduced eective values for CA in a nuclear medium, and spin-spin
correlations which tend to pair the spins and thus tend to reduce the
opacities. The question of the appropriate neutrino opacities in a SN
medium remains worrisome.
However, with regard to particle bounds the eect of reduced opacities goes in the direction of making those constraints more conservative
as a reduction of the opacities, like an anomalous energy loss, shortens
the neutrino signal.
13.8
Right-Handed Neutrinos
13.8.1
Dirac Mass
Right-handed neutrinos (helicity-minus neutrinos, helicity-plus antineutrinos) do not interact by the standard weak interactions and so
they would not be trapped in the interior of a SN core. Therefore, the
SN 1987A neutrino signal allows one to constrain any mechanism that
517
0.71019 erg g1 s1
m
30 keV
)2 (
T
30 MeV
)4
(13.14)
where CV2 +3CA2 1 was used (see Appendix B). Even though the axialvector structure function is not a function, Eq. (13.14) is probably a
reasonable estimate because the results of the previous section indicate
that the neutrino scattering rate in a dense medium is probably not
86
Such bounds naturally can be avoided if one assumes that the r.h. neutrinos
have other novel interactions which are strong enough to trap them eciently in a
SN core. Explicit models were constructed, for example, by Babu, Mohapatra, and
Rothstein (1992) or Rajpoot (1993). These authors aimed at avoiding the SN 1987A
bound on Dirac neutrino masses.
87
Besides the neutrino spin-ip rate due to the neutrino weak interactions with
nucleons or other particles, there is also a spin-ip scattering term in the gravitational eld of the entire neutron star (Choudhury, Hari Dass, and Murthy 1989).
However, the resulting energy loss was found to be small except for low-energy
neutrinos.
518
Chapter 13
too dierent from that found in the dilute-medium limit. If one applies
the analytic criterion Eq. (13.8) one nds
m <
30 keV
(13.15)
519
neutrino signal is not aected as signicantly as one might have expected. The Dirac mass limit becomes only slightly more restrictive if
mixing is assumed (Burrows, Gandhi, and Turner 1992).
Neutrinos with masses in the keV range must decay suciently fast
in order to avoid overclosing the universe. If r.h. Dirac-mass neutrinos escape directly from the inner core of a SN they have energies
far in excess of l.h. neutrinos emitted from the neutrino sphere. If
their decay products involve sequential l.h. neutrinos or antineutrinos,
these daughter states would have caused high-energy events at IMB or
Kamiokande II, contrary to the observations. Dodelson, Frieman, and
Turner (1992) found that this argument excludes the lifetime range
7
<
109 s/keV <
/m 510 s/keV,
(13.16)
<
for Dirac masses in the range 1 keV <
m 300 keV, assuming that
the visible channel dominates.
13.8.2
Right-Handed Currents
On some level r.h. weak gauge interactions may exist as, e.g. in leftright symmetric models where the gauge bosons which couple to r.h.
currents would dier from the standard ones only in their mass. In the
low-energy limit relevant for processes in stars one may account for the
novel couplings by a r.h. Fermi constant which is given as GF with
some small dimensionless number which may be dierent for chargedand neutral-current processes. In left-right symmetric models one nds
explicitly for charged-current reactions (Barbieri and Mohapatra 1989)
2CC = 2 + (mWL /mWR )4 ,
(13.17)
where mWR,L are the r.h. and l.h. charged gauge boson masses while
is the left-right mixing parameter.
In order to constrain CC one assumes the existence of r.h. e s so
that the dominant energy-loss mechanism of a SN core is e + p
n + e,R where the nal-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 + e,R . The usual
charged-current weak scattering cross section involving nonrelativistic
nucleons is G2F (CV2 +3CA2 )Ee2 / with CV2 +3CA2 4 in a nuclear medium
(Appendix B). Using a proton density corresponding to nuclear matter
at 1015 g cm3 and using 100 MeV for a typical electron energy one nds
520
Chapter 13
(13.18)
in agreement with a result of Barbieri and Mohapatra (1989) while Raffelt and Seckel (1988) found a somewhat less restrictive limit of CC <
2
310
3105 . Laboratory experiments yield a limit of order CC <
(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 e .
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 R 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 R 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 (15 = 0.03) and at T = 30 MeV, and applying the analytic
criterion Eq. (13.8) one nds
3
NC <
310 .
(13.19)
This is less restrictive from what was found by Raelt 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. Detailed analyses were presented by Grifols and Masso (1990b), Grifols,
Masso, and Rizzo (1990), and Rizzo (1991). Because these authors did
not consider multiple-scattering eects 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
3
from big bang nucleosynthesis (NC <
10 ) which are based on the
requirement that r.h. neutrinos must not have come to thermal equilibrium after the QCD phase transition (at T <
200 MeV) in the early
universe (e.g. Olive, Schramm, and Steigman 1981; Ellis et al. 1986).
521
(13.20)
522
Chapter 13
As the r.h. neutrinos escape from the inner core with much larger energies than those from the neutrino sphere one would expect high-energy
events in the Kamiokande and IMB detectors, contrary to the observa12
tions. Therefore, one probably needs to require <
10 B for the
diagonal dipole moments; spin-avor oscillations could be suppressed
by the neutrino mass dierences. 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 1012 B could also aect the
infall phase of SNe. The spin-ip scattering on nuclei would be coherently enhanced relative to protons. Therefore, neutrinos could escape
in the r.h. channel for much longer so that eectively trapping would
set in much later than in the standard picture (Notzold 1988).
In and near the SN core there probably exist strong magnetic elds
of order 1012 Gauss or more which would induce spin-precessions between r.h. and l.h. neutrinos. Therefore, the sterile states produced in
the deep interior by spin-ip scattering could back-convert into active
ones near the neutrino sphere. Depending on details of the matterinduced neutrino energy shifts, the vacuum mass dierences, and the
magnetic eld strengths and congurations this conversion could take
place inside or outside of the neutrino sphere. The observable neutrino
signal could be aected, but also the energy transfer within the SN core
and outside of the neutrino sphere. Perhaps, a more ecient transfer of
energy to the stalled shock wave could help to explode SNe in the delayed 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 1012 B range
and below would aect SN dynamics and the observable neutrino signal
in interesting ways. However, because there are so many parameters
and possible eld congurations, 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 impossible 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
523
(13.21)
Charge Radius
Chapter 14
Axions
The idea of axions is introduced and their phenomenological properties
are reviewed. The constraints on pseudoscalars that have been derived
throughout this book are systematically applied to axions.
14.1
524
Axions
525
Table 14.1. Particle magnetic and electric dipole moments.
Fermion
Protona
Neutrona
Electrona
Neutrinob
Magnetic Momentc
Electric Momentd
[1026 e cm]
4000 6000
< 11e
0.3 0.8
< 6000
a Particle
s e
GG .
8
(14.1)
e
Here, s is the ne-structure constant of strong interactions and GG
e
1
e
Gb Gb where Gb is the color eld strength tensor, G
b = 2 Gb
its dual, and the implied summation over b refers to the color degrees
of freedom. Of course, det Mq and thus would vanish if one of the
quarks were exactly massless, but this does not seem to be the case.
Under the combined action of charge conjugation (C) and a parity
transformation (P) the Lagrangian Eq. (14.1) changes sign,89 violating
the CP invariance of QCD. It leads to a neutron electric dipole moment
|dn | || (0.04 2.0)1015 e cm (Baluni 1979; Crewther et al. 1979;
see also Cheng 1988). This is |dn | || (0.004 0.2) N in units of
nuclear magnetons. Hence, for || of order unity one expects a neutron electric dipole moment almost as large as its magnetic one. The
e is Ecolor Bcolor , i.e. the scalar product of a polar with an
The structure of GG
axial vector and so it is CP-odd.
89
526
Chapter 14
9
experimental limit (Tab. 14.1), however, indicates || <
10 , surpris3
ingly small in view of the phase = 3.310 which appears in the
Cabbibo-Kobayashi-Maskawa matrix Eq. (7.6) and which explains the
observed CP-violating eects in the K -K system.
Even worse, QCD alone produces a term like Eq. (14.1) because of
the nontrivial topological structure of its ground state (Callan, Dashen,
and Gross 1976; Jackiw and Rebbi 1976; tHooft 1976a,b). The coecient QCD is a parameter characterizing the -vacuum. It is mapped
onto itself by a transformation QCD QCD + 2 so that dierent
ground states are characterized by values in the range 0 QCD < 2.
The phase of the quark mass matrix and the QCD-vacuum together
yield QCD + arg det Mq as a compound coecient for Eq. (14.1).
The experimental bounds then translate into
9
QCD + arg det Mq <
10 .
(14.2)
14.2
14.2.1
Generic Features
In the literature one often nds fa /N , Fa /N , vPQ /N etc. for what I call fa . It
was stressed, e.g. by Georgi, Kaplan, and Randall (1986) that a discussion of the
generic properties of all axion models does not require a specication of the modeldependent integer N which can be conveniently absorbed in the denition of fa .
Axions
527
s
e.
a GG
8 fa
(14.3)
(14.4)
where m = 135 MeV is the pion mass and f 93 MeV its decay
constant. This mass term implies that at low energies the axion Lagrangian contains a potential V (a) which expands to lowest order as
1 2 2
m a . Because of the invariance of L with respect to + 2
2 a
the gluon-induced potential V (a) is a function periodic with 2fa .
Fig. 14.1. Axion mixing with qq states and thus . The curly lines represent
gluons, the solid lines quarks.
The ground state of the axion eld is at the minimum of its potential at a = 0, explaining the absence of a neutron electric dipole
moment. If one could produce a static nonvanishing axion eld a0 in
some region of space, neutrons there would exhibit an electric dipole
moment corresponding to = a0 /fa .
The Lagrangian Eq. (14.3) is the minimal ingredient for any axe coupling is their dening feature as opposed to
ion model: the aGG
other pseudoscalar particles. Then axions inevitably acquire an eective 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 , axions share not only their mass, but
also their couplings to photons and nucleons with a strength reduced
by about f /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 eective axion mass is a low-energy phenomenon below QCD
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 < 2fa 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 dierent initial
conditions in dierent regions of space. As the universe expands and
cools below QCD , however, the axion eld 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),
a 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
5
the universe. With Eq. (14.4) this corresponds to ma <
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.
14.2.2
i
2
+ h.c. + V (||)
(
h L R + h.c. .
(14.6)
Axions
529
L ei/2 L ,
R ei/2 R ,
(14.7)
where the left- and right-handed elds pick up opposite phases. This
chiral symmetry is usually referred to as the Peccei-Quinn (PQ) symmetry UPQ (1).
The potential V (||) is chosen
to be a Mexican hat with an absolute minimum at || = fPQ / 2 where fPQ is some large energy scale.
The ground state is characterized
by a nonvanishing vacuum expecta i
tion value = (fPQ / 2) e where is an arbitrary phase. It spontaneously breaks the PQ symmetry because it is not invariant under a
transformation of the type Eq. (14.7). One may then write
=
fPQ + ia/fPQ
e
2
(14.8)
in terms of two real elds and a which represent the radial and
angular excitations.
The potential V provides a large mass for , a eld which will be of
no further interest for these low-energy considerations. Neglecting all
terms involving the Lagrangian Eq. (14.6) is
L=
530
Chapter 14
of a with gluons is then given by the triangle graph of Fig. 14.2. With
the rst term of Eq. (14.10) it yields an eective a-gluon interaction of
LaG =
ga s
e,
a GG
m 8
(14.11)
where s gs2 /4. 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 analogous graph pertains to the coupling of axions with photons if the fermion
carries an electric charge which replaces gs .
(14.12)
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 eld.
The potential V (a) is periodic with 2fa = 2fPQ /N . The interpretation of a as the phase of , on the other hand, implies a periodicity
with 2fPQ so that N must be a nonzero integer. This requirement
restricts the possible assignment of PQ charges to the quark elds. It
also implies that there remain N dierent equivalent ground states for
the axion eld, each of which satises = 0 and thus solves the CP
problem.
Axions
14.2.3
531
Pseudoscalar vs. Derivative Interaction
R eia/2fPQ R .
(14.14)
The last term in Eq. (14.10) is then a simple mass term m. The
interaction between and a now arises from the kinetic term in
Eq. (14.9),
Lint =
1
5 a .
2fPQ
(14.15)
(14.16)
532
Chapter 14
with the axion example Eq. (14.15) where only one Nambu-Goldstone
boson was present as opposed to the pion isotriplet. However, this additional 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 (Raelt 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 e e a one may use either the pseudoscalar 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.
14.2.4
A heavy critique was levied against the PQ mechanism by quantumgravity inspired phenomenological considerations. The main idea is
that generally the PQ symmetry, like any other global symmetry, will
not be respected by gravity (Georgi, Hall, and Wise 1981). For example, a black hole can swallow any amount of PQ charge without a
trace, while a swallowed electric charge remains visible by its Coulomb
force. At energy scales exceeding the Planck mass mPl = 1.21019 GeV
quantum gravitational eects are expected and so mPl is a phenomenological cuto for any quantum theory which does not fundamentally
include gravitation. In the low-energy world it should manifest itself
by all sorts of eective interactions which are not forbidden by a symmetry and which likely involve inverse powers of the cuto scale mPl .
Axions
533
Notably, the Higgs eld which gives rise to the axion probably
exhibits eective interactions of dimension 2m + n
m n
i ( )
Vgrav () = g e
,
(14.17)
m2m+n4
Pl
where g and are real numbers. Because such interactions violate the
PQ symmetry for n = 0 they induce an eective 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
1
cos(
+
na/f
)
,(14.18)
1
cos(a/f
)
+
m
=
m
a
a
QCD
grav
fa2
where mQCD is the usual QCD axion mass while gravity induces
For g of order unity and for low values of m and n one needs a very
small fa for the QCD eect 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 eects 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 NambuGoldstone boson. These particles are interlopers in the low-energy
worldaxions 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 experimental 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 Sha (1986), and Dine and Seiberg (1986).
534
Chapter 14
14.3
14.3.1
Axions generically mix with pions so that their mass and their couplings
to photons and nucleons are crudely f /fa times those of . In detail,
however, these properties depend on the specic implementation of
the PQ mechanism. Therefore, it is useful to review briey 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 eld of which axions are the phase to be the standard Higgs
eld. Therefore, one needs to introduce two independent
Higgs elds 1
2
and
f
/
2 which must
and 2 with vacuum expectation
values
f
/
2
1
2
1/2
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)
1 gives masses to the up- and 2 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 anelectroweak singlet Higgs eld
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 referred to as invisible axion models. The rst 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 Zhitnitski
(1980) and by Dine, Fischler, and Srednicki (1981). It is a hybrid between the standard and KSVZ models in that it uses an
electroweak
singlet scalar eld with a vacuum expectation value fPQ / 2 and two
electroweak doublet elds 1 and 2 . 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 or equivalently by
cos2 = x2 /(x2 + 1).
From a practical perspective, the main dierence 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 unication scale fGUT 1016 GeV (Wise, Georgi,
and Glashow 1981; Nilles and Raby 1982). However, the cosmologi12
cal bound fa <
10 GeV disfavors the GUT assignment. There exist
numerous other axion models, and many attempts to connect the PQ
scale with other scalesfor a review see Kim (1987). In the absence
of a compelling model fa should be viewed as a free phenomenological
parameter.
14.3.2
The axion mass which arises from its mixing with 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)
f m
ma =
fa
z
(1 + z + w)(1 + z)
107 GeV
= 0.60 eV
,
fa
)1/2
(14.20)
where the quark mass ratios are (Gasser and Leutwyler 1982)
z mu /md = 0.568 0.042,
w mu /ms = 0.0290 0.0043.
(14.21)
536
Chapter 14
(14.22)
Xj Q2j Dj ,
(14.23)
where Dj = 3 for color triplets (quarks) and 1 for color singlets (charged
leptons). The total axion-photon coupling strength is then (Kaplan
1985; Srednicki 1985)
3
meV
ga =
=
,
(14.24)
2fa 4
0.691010 GeV
where
(
)
(
)
4 E
2 4+z+w
4 E
=
1.92 0.08
(14.25)
3 N
3 1+z+w
3 N
and meV ma /eV.
In the DFSZ or grand unied models one has for a given family of quarks and leptons E/N = 8/3. Neglecting w this yields
(8/3) z/(1 + z) 1. However, one may equally consider models where
E/N = 2 so that = 0.10.1, i.e. the axion-photon coupling is strongly
suppressed and may actually vanish (Kaplan 1985).
14.3.3
The discussion in Sect. 14.2.3 implies that axions interact with a given
fermion j (mass mj ) according to a pseudoscalar or a derivative axialvector interaction,
Cj mj
Cj
Lint = i
j 5 j a or
j 5 j a ,
(14.26)
fa
2fa
where Cj is an eective PQ charge of order unity to be dened below.
Evidently gaj Cj mj /fa plays the role of a Yukawa coupling and
2
/4 that of an axionic ne structure constant. Numerically,
aj = gaj
gae = Ce me /fa = Ce 0.851010 meV ,
gaN = CN mN /fa = CN 1.56107 meV
for electrons and nucleons.
(14.27)
Axions
537
(14.28)
(14.29)
d = 0.41,
s = 0.08
(14.30)
538
Chapter 14
(14.31)
(14.32)
In Fig. 14.4 these couplings are shown, for DFSZ axions as a function
of cos2 . They are all uncertain to within about 0.05, but even then
Cp and Cn never seem to vanish simultaneously.
14.4
Barr and Seckel (1992) studied astrophysical axion bounds when quantum gravity eects are taken seriously.
Axions
539
Bounds on the Yukawa coupling to electrons of various novel particles were derived in Chapter 3; for pseudoscalars a summary was given
in Tab. 3.1. The most restrictive limit was obtained from the delay of
helium ignition in low-mass red giants that would be caused by excessive axion emission; in terms of the axion-electron Yukawa coupling it
13
is gae <
2.510 . With Eq. (14.27) this translates into
9
>
ma Ce <
0.003 eV and fa /Ce 210 GeV.
(14.33)
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 nds 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 nds
2
9
>
ma cos2 <
0.01 eV and fa / cos 0.710 GeV.
(14.34)
(14.35)
(14.36)
540
Chapter 14
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 essentially do not couple to neutrons while Cp 0.36. For DFSZ axions
the couplings vary with cos2 , although for cos2 0.5 about the same
values as for KSVZ axions apply which are thus taken as generic. Assuming a proton fraction of about 0.3 for the relevant regions of the SN
core I estimate an eective nucleon coupling of CN 0.31/2 0.36 0.2.
Therefore,
<
0.01 eV <
ma 10 eV,
9
<
0.6106 GeV <
fa 0.610 GeV
(14.37)
14.5
Cosmological Limits
Axions
541
542
Chapter 14
(14.38)
Axions
543
544
Chapter 14
Chapter 15
Miscellaneous Exotica
Stellar-evolution constraints on a variety of hypotheses are discussed
and compared with limits from other sources. Specically, a possible
time variation of Fermis and Newtons constant, the validity of the
equivalence principle, a photon mass and charge, the existence of free
quarks and supersymmetric particles, and the role of majorons and
millicharged particles are considered.
15.1
One of the basic physical assumptions commonly made in astrophysical research is that the laws of nature are the same at dierent places
in the universe, and at earlier times here and elsewhere. Apparently
this assumption has never been challenged seriously by any experiment
or observation that would have indicated a spatial or temporal variation of parameters such as particle masses or coupling constants. At
the present time it is not known what xes the values of such fundamental numbers. Therefore, the possibility that they vary in time or
space cannot be a priori rejected. Notably, Dirac (1937, 1938) is often
quoted for his speculation that the large value of some dimensionless
numbers occurring in physics are related to variations of some physical
constants on cosmological time scales. Whatever the merit of Diracs
large numbers hypothesis, it remains an interesting task to isolate simple observables that are sensitive to variations of certain constants.
One instructive stellar-evolution example was discussed by Scherrer
and Spergel (1993) who considered the constancy of Fermis constant
GF which governs weak-interaction
2 physics. The particle-physics stan1
dard model gives GF = 2 0 in terms of the Higgs-eld vacuum
expectation value 0 . Scherrer and Spergel noted that a nonconstant
545
546
Chapter 15
15.2
15.2.1
Another constant of nature that might vary in time is Newtons constant GN . Indeed, there exist self-consistent alternative theories to
general relativity which actually predict a temporal variation of GN on
cosmological time scalessee Will (1993) for a summary of such theories and detailed references. A typical scale for the rate of change
is the cosmic expansion parameter H so that it is natural to write
G N /GN = H with a dimensionless model-dependent number. Some
93
The Earth moves with the galaxy and the local group relative to the cosmic
microwave background (CMB). Taking 600 km s1 for this peculiar velocity the
Earth moves by about 1.2 Mpc in 2 Gyr relative to a frame dened by the CMB.
Miscellaneous Exotica
547
Table 15.1. Bounds on the present-day G N /GN . (Adapted from Will 1993.)
Method
G N /GN
[1012 yr1 ]
References
M
uller et al. (1991)
Shapiro (1990)
Damour and Taylor (1991)
Goldman (1990)
Big-Bang Nucleosynthesis
GN
the universe the expansion rate is given by H 2 = (R/R)
= 8
3
in terms of the energy density which, during the epoch of nucleosynthesis, is dominated by radiation (photons, neutrinos). It is a standard
94
548
Chapter 15
(15.1)
where t0 refers to the present epoch. Assuming that GN at nucleosynthesis was within 50% of its standard value one nds || <
0.01 which
12
1
would imply |G N /GN |today <
10
yr
,
at
least
a
factor
of ten below
the present-day limits. One should keep in mind, however, that a powerlaw variation of GN is a relatively arbitrary assumption. For example,
in scalar-tensor extensions of general relativity such as the Brans-Dicke
theory GN varies as a power law during the matter-dominated epoch
while it remains constant when radiation dominates. Either way, while
the nucleosynthesis bounds are probably somewhat more restrictive
than the celestial-mechanics ones it is interesting that the resulting
12
1
bounds |G N /GN |today <
11010 yr are of the same general order
of magnitude. They leave room for a considerable variation of GN over
cosmic time scales.
Miscellaneous Exotica
15.2.3
549
550
Chapter 15
Table 15.2. Characteristics of the Demarque et al. (1994) solar models with
a varying GN according to Eq. (15.1).
Xinitial
[%]
Xc
[%]
Tc
[106 K]
c
[g/cm3 ]
Renv
[R ]
15.14
15.28
15.37
15.47
15.58
15.72
16.07
125.5
134.0
139.5
146.2
154.6
165.1
197.3
0.739
0.731
0.724
0.721
0.716
0.710
0.695
37
Cl 71 Ga
[SNU] [SNU]
5.5
117
6.8
124
8.7
134
(15.2)
similar to the celestial-mechanics bounds of Tab. 15.1. The precise functional form Eq. (15.1) is not crucial for the solar bound as it probes GN
only for the last 4.5 Gyr of the assumed 15 Gyr cosmic age. Therefore,
one could have equally assumed a linear form for GN (t).
Miscellaneous Exotica
551
White Dwarfs
Globular Clusters
The oldest stellar objects in the galaxy are globular-cluster stars which
are thus expected to yield the most restrictive stellar-evolution limits
on G N /GN . A color-magnitude diagram for an intermediate-aged galactic cluster was constructed by Roeder (1967) while detailed studies of
globular clusters were performed by Prather (1976) and DeglInnocenti
et al. (1995). Roeder (1967) and Prather (1976) used a time variation
for a specic Brans-Dicke cosmology where GN decreases approximately
as in Eq. (15.1) with 0.03 while DeglInnocenti et al. (1995) considered more generic cases of GN (t).
552
Chapter 15
t0
t0
dt GN (t)/GN (t0 )
(15.3)
GN (t) = 1 + 0 (t t0 ) GN (t0 ),
(15.4)
=
=
,
1 (1 0 )1
0
(15.5)
Miscellaneous Exotica
553
Fig. 15.1. Required present-day G N /GN in order to achieve a true globularcluster age , given that the apparent age is . A linear GN (t) variation as
in Eq. (15.4) was assumed.
can relate a desired true age to a required value for 0 (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
12
1
G N /GN <
710 yr .
(15.6)
A lower age limit is less certain. Taking 8 Gyr one nds G N /GN >
351012 yr1 which is less certain, and also less interesting relative
to the limits discussed in the previous sections.
554
Chapter 15
Actually, a certain reduction of the true globular-cluster ages relative to their apparent ones would be a welcome cosmological eect as
they are, at best, marginally compatible with other cosmic age indicators. In view of the current limits on G N /GN summarized in Fig. 15.2
this possibility cannot be excluded at present.
15.3
The equivalence principle of Einsteins general theory of relativity implies that the space-time trajectories of relativistic particles should be
independent of internal degrees of freedom such as spin or avor, and
independent of the type of particle under consideration (photons, neutrinos). A number of astronomical observations allow one to test this
prediction. Laboratory tests of various consequences of the equivalence
principle are discussed in Wills (1993) book.
Nonsymmetric extensions of general relativity (e.g. Moat 1991)
predict that dierent polarization components of electromagnetic waves
propagate with dierent phase velocities in gravitational elds. This
birefringence eect would lead to the depolarization of the Zeeman
components of spectral lines emitted in magnetically active regions of
the Sun. The absence of this depolarization eect leads to signicant
constraints on Moats theory and others (Gabriel et al. 1991).
In a similar approach one uses the dierence of the Shapiro time
delay between dierent particles or between dierent polarization states
of a given particle which propagate through the same gravitational
eld. In Sect. 13.3 the absence of an anomalous shift between the
SN 1987A photon and neutrino arrival times gave limits on violations
of the equivalence principle because both pulses moved through the
same galactic gravitational potential.
Also, one may search for dierences in the arrival times of leftand right-handed polarized electromagnetic signals from distant pulsars
(LoSecco et al. 1989). The best bound was obtained from an analysis of
the pulse arrival times from PSR 1937+21 which is about 2.5 kpc away
from Earth. One may write the eective gravitational potential in the
form V (r) = V0 (r) [1 + A1 r + A2 v + A3r (v )] where r, v, and
represent the location, velocity, and spin of the particles (photons,
neutrinos). The PSR 1937+21 data then yield a constraint |A1 | <
41012 and |A2 | < 11012 (Klein and Thorsett 1990), apparently
the most restrictive limits of their kind.
Miscellaneous Exotica
555
15.4
556
Chapter 15
charged particles in the galactic magnetic eld would be curved, leading to an energy-dependent time-delay (Barbiellini and Cocconi 1987).
Because the same argument can be applied to photons, the signals from
radio pulsars also allow one to set a limit on a putative photon electric
charge (Cocconi 1988). However, the resulting dispersion eect scales
with photon frequency in the same way as the eect caused by a photon
mass or by the plasma eect so that this method, again, is limited by
the standard dispersion eect (Raelt 1994). One nds a bound on the
29
photon charge of Q <
10 e.
Returning to a hypothetical photon mass, its value can be extracted,
in principle, from the spatial distribution of static magnetic elds of
celestial bodies. The measured elds can be tted by an appropriate
multipole expansion in which m is kept as a free parameter. The
most restricitve limit of this sort was derived from Jupiters magnetic
eld on the basis of the Pioneer-10 observations; Davis, Goldhaber,
15
and Nieto (1975) found a limit m <
0.610 eV. The same method
applied to the Earths magnetic eld yields an almost equivalent bound
15
of m <
0.810 eV (Fischbach et al. 1994).
As detailed in a review by Barrows and Burman (1984) more restrictive limits obtain from detailed considerations of astrophysical objects
in which magnetic elds, and hence the Maxwellian form of electrodynamics, play a key role in maintaining equilibrium or creating long-lived
27
stable structures. The most restrictive such limit of m <
10 eV is
based on an argument by Chibisov (1976) concerning the magnetogravitational equilibrium of the gas in the Small Magellanic Cloud
which requires that the range of the interaction exceeds the characteristic eld scale of about 3 kpc. This limit, if correct, is surprisingly
close to 1033 eV where the photon Compton wavelength would exceed
the radius of the observable universe and thus would cease to have any
observable consequences.
15.5
Free Quarks
Miscellaneous Exotica
557
15.6
Supersymmetric Particles
558
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, Masso, and
Peris 1989; Grifols and Masso 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. Moreover, 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 particle 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
Axions (Chapter 14) are one representative of a variety of NambuGoldstone bosons of spontaneously broken global symmetries that have
appeared in the literature over the years. Another widely discussed example are the majorons rst 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 neutrinos 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 conjecture 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 interactions 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 signicant
Miscellaneous Exotica
559
560
Chapter 15
masses by a see-saw type mixing eect with the heavy states. The majoron coupling to standard neutrinos would be extremely small in this
model, leading to no interesting consequences besides small Majorana
masses for e , , and .
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 eld. 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
eld to fermions scale with the inverse of v, the majoron would have
a rather strong coupling to neutrinos. Among many fascinating phenomenological and astrophysical consequences (e.g. Georgi, Glashow,
and Nussinov 1981; Gelmini, Nussinov, and Roncadelli 1982) this model
predicted, however, that the new Higgs eld 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 19891990,
however, correspond exactly to the known three neutrino avors (Particle Data Group 1994), leaving no room for this triplet majoron model.
Other doublet majoron models which would contribute one-half of an
eective 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. Berezhiani, 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 ones way to construct such
models does not arise from particle theory but rather from experiments 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 2 e ) can also decay in a neutrinoless mode if e has a Majorana
mass, allowing an emitted e to be eectively reabsorbed as a e . In
majoron models of Majorana neutrino masses there is a third decay
channel where the intermediate e in the 0 mode radiates a majoron
so that eectively 2e plus one majoron are emitted. The expected
sum spectrum of the electron energies would be continuous as in the
2 mode, but with a dierent spectral shape. Once in a while, experiments which search for the 0 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 2 mode or other
Miscellaneous Exotica
561
562
Chapter 15
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 2 experiments do not yield any compelling evidence for majoron emission, one may consider a decayingneutrino cosmology as a motivation in its own right for majoron models. Such cosmologies may explain the discrepancy between the cosmic
density uctuation spectrum inferred from the cosmic microwave background 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 neutrinorelated 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 couplings to neutrinos (the Gelmini-Roncadelli model) is experimentally
excluded although one can construct more complicated ones which retain sizeable neutrino-majoron couplings and yet are compatible with
the Z decay width. The possibility of such models is entertained because 2 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.
15.7.2
Majorons could also have an impact on stellar evolution. Besides interacting with neutrinos, they typically also couple to other fermions,
allowing one to apply the astrophysical bounds on pseudoscalars derived throughout this book to majoron models.
Because of the possibility of fast decays the neutrino signal
from distant sources, notably from the Sun or from SN 1987A, would
be aected. Such decays are not likely to be able to explain the solar neutrino problem as discussed in Sect. 10.8. It remains interesting,
however, that the matter-induced e - e energy splitting allows for decays e e (Sect. 6.8). Should a solar e ux show up in future
measurements, it could be an indication for such decays.
As for SN neutrinos, the decay of massive s or s with nalstate e s could modify the e signal observed in a detector, notably
the energy distribution and duration of the observed pulse (Soares and
Wolfenstein 1989; Simpson 1991). The SN 1987A data have not been
analyzed in detail with regard to this possibility, although the observed
signal can be accounted for without invoking such eects. It is not
clear if one could derive signicant constraints on majoron models from
Miscellaneous Exotica
563
Kolb, Tubbs, and Dicus (1982); Dicus, Kolb, and Tubbs (1983); Manohar
(1987); Fuller, Mayle, and Wilson (1988); Aharonov, Avignone, and Nussinov
(1988b, 1989); Choi et al. (1988); Grifols, Masso, and Peris (1988); Konoplich
and Khlopov (1988); Dicus et al. (1989); Berezhiani and Smirnov (1989); Choi and
Santamaria (1990).
564
Chapter 15
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
15.8
Millicharged Particles
Miscellaneous Exotica
565
(15.7)
(15.8)
(15.9)
(15.10)
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 restrictive bounds from a host of accelerator experiments (Fig. 15.3).
A simple astrophysical constraint is based on avoiding excessive energy losses of stars which can produce millicharged particles by various
reactions, most notably the plasma decay process. To avoid an unacceptable delay of helium ignition in low-mass red giants, and to avoid an
566
Chapter 15
Miscellaneous Exotica
567
Chapter 16
Neutrinos: The Bottom Line
Most of the particle-physics arguments discussed in this book are closely
related to neutrino physics because these particles play an important
role in stellar evolution whether or not they have nonstandard properties. Besides a summary of some recent developments of standardneutrino astrophysics, a synthesis is attempted of what stars as neutrino
laboratories have taught us about these elusive objects, and what one
might reasonable hope to learn in the foreseeable future.
16.1
Standard Neutrinos
The main theme of this book has been an attempt to extract information about the properties of neutrinos and other weakly interacting particles from the established properties of stars. However, even
standard-model neutrinos (massless, no mixing, no exotic properties)
play a signicant role in astrophysics. There have been some recent developments in standard-neutrino astrophysics which deserve mention
in a summary.
It is now thought that neutrinos play an active role in supernovae
besides carrying away the binding energy of the newborn neutron star
(Chapter 11). In the delayed-explosion scenario they are crucial to
revive the stalled shock wave which is supposed to expel the stellar
mantle and envelope. Moreover, they have a strong impact on r-process
nucleosynthesis which is thought to occur in the high-entropy region
above the neutron star a few seconds after collapse. For both purposes it
is crucial to calculate the SN neutrino lightcurve for the rst seconds
after collapse. Convection below the neutrino sphere and large-scale
convective turnovers in the region between the neutron star and the
shock wave are both important and need to be understood better on
568
569
570
Chapter 16
16.2
16.2.1
Neutrinos with nonstandard properties would have more radical implications in astrophysics. A minimal and most plausible extension of the
standard model is the possibility that they have masses and mixings
like the other fermions. Taking this hypothesis in a literal sense means
that neutrinos would need to have Dirac masses like the charged fermions. Therefore, one needs to postulate the existence of right-handed
neutrinos and Yukawa couplings to the standard Higgs eld to generate
masses and mixings. What do we know about neutrinos in the context
of this Minimally Extended Standard Model?
Perhaps the most dramatic lesson is that the mass of all sequential neutrinos (e , , ) must be less than the cosmological limit of
approximately 30 eV (Sect. 7.1.5). This conclusion is not to be taken
for granted as massive neutrinos with mixings can decay by virtue of
, and by e e+ e if m exceeds about 2me . While the
standard-model radiative decays are too slow to avoid the cosmological
limit, the e+ e channel can be fast on cosmological time scales if the
mixing angle is not too small. This decay channel is an option only
for which may have a mass of up to 24 MeV while the experimental
mass limits on the other avors are below 0.16 MeV.
Such a heavy , however, can be excluded by several arguments.
The stellar evolution one based on SN arguments was presented in
Sect. 12.5.2. The bremsstrahlung emission of photons in e e+ e
would produce a -ray ux in excess of the SMM limits for SN 1987A
9
unless sin2 2e3 <
10 . (For a detailed dependence of this limit on the
assumed mass see Fig. 12.18.) In addition, the integrated positron ux
from all galactic supernovae over the past, say, 100,000 years would
exceed the observed value unless the decays are very fast (near the SN)
or very slow (outside of the galactic disk). Together, these limits leave
no room for a heavy (Fig. 12.19). In addition, the mass range
<
0.5 MeV <
m 35 MeV can be excluded on the basis of big bang
nucleosynthesis arguments (Fig. 7.2) unless the neutrinos are shorter
lived than permitted by the Minimally Extended Standard Model.
Neutrino masses near the cosmological limit would be important for
cosmology as they could contribute some or all of the dark matter of
the universe. The latter option is disfavored by theories of structure
formation. However, a subdominant hot dark matter contribution in
the form of, say, 5 eV neutrinos might be cosmologically quite welcome.
571
Neutrino masses in this general range are also accessible to timeof-ight measurements. The neutrino burst of SN 1987A has already
provided a limit of me <
20 eV (Sect. 11.3.4), a result which remains
interesting in view of the confusing situation with the tritium decay endpoint experiments which seem to be plagued by systematic effects which cause the appearance of a negative neutrino mass-square
(Sect. 7.1.3). The observation of the prompt e burst from a future
galactic SN would allow one to reduce this limit to a few eV. Moreover, one may well be able to detect or constrain a or mass in
the 1020 eV range, assuming the simultaneous operation of a water
Cherenkov detector such as Superkamiokande and a neutral-current detector such as the proposed Supernova Burst Observatory (Sect. 11.6).
16.2.2
Solution
m2 [eV2 ]
sin2 2
Large-angle MSW
Nonadiabatic MSW
Vacuum oscillations
2105
0.6105
0.81010
0.6
0.006
0.81
572
Chapter 16
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 uxes in experiments with three dierent
spectral response characteristics. Even allowing for large modications
of the p 7 Be cross section or the solar central temperature does not yield
consistency unless one stretches the experimental uncertainties of the
ux measurements beyond reasonable limits. Still, a nal 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.
16.2.3
573
Electromagnetic Properties
574
Chapter 16
16.3
New Interactions
16.3.1
Majorana Masses
575
576
Chapter 16
state general constraints as one may easily postulate, for example, that
the decays do not involve nal-state e s, that even wrong-helicity
Dirac neutrinos are trapped in a SN core by novel interactions, or that
heavy Majorana s have only negligible mixings with e . Surely other
loopholes could be found.
16.3.3
Electromagnetic Properties
577
of temperature and lepton number. (Of course, a nonlocal energy transfer mechanism is already thought to be important for reviving the shock
wave in the delayed-explosion scenario.) As far as I know, nothing more
quantitative than back-of-the-envelope estimates of this scenario exist
in the literature. Therefore, alleged bounds of order 1012 B (Bohr
magneton B = e/2me ) on Dirac-neutrino dipole moments probably
have to be used with some reservation (Sect. 13.8.3). Conversely, such
dipole moments may actually help to explode supernovae.
With regard to the observable neutrino signal, spin and spin-avor
oscillations both in the SN and in the galactic magnetic eld may cause
vast modications of the e uxes and spectra observable in water Cherenkov detectors if neutrinos have dipole moments in the ballpark of
1012 1014 B .
Spin and spin-avor oscillations can be very important in the early
universe where strong magnetic elds may exist, and where a population of the r.h. degrees of freedom would accelerate the expansion
rate of the universe. These issues are being investigated in the current
literature; nal conclusions do not seem to be available at the present
time. Still, it appears that this eect may well be the most signicant
impact of small Dirac neutrino dipole or transition moments anywhere
in nature.
b) Laboratory Limits
Less problematic bounds on neutrino dipole moments arise from laboratory experiments where one studies the recoil spectrum of electrons
in the reaction + e e + where can be the same or a dierent
avor (Sect. 7.5.1). A sensitivity down to, perhaps, as low as 1011 B
can be expected from a current eort involving reactor neutrinos as
a source (MUNU experiment). Current limits on dipole or transition
moments are about 21010 B if e is involved, and about 71010 B
if 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.
c) Huge Dipole Moments or Millicharges
In principle, the possibility of a large diagonal moment for remains
open as it has not been possible to produce a strong source in the laboratory so that only extremely crude limits exist on the -e-scattering
cross section. The globular-cluster bounds discussed below do not
578
Chapter 16
apply to a heavy so that one is confronted with a nontrivial allowed region in -m space where even MeV masses become cosmologically allowed because of the dipole-induced annihilation process
e+ e ( in the ballpark of 107 B ). Of course, such large
dipole moments must be caused by a fairly nontrivial arrangement of
intermediate charged states and so one may wonder if a large magnetic moment could be realistically the only manifestation of these new
particles and/or interactions.
Still, a large dipole moment and the correspondingly large annihilation cross section in the early universe is one possibility to tolerate a large mass without the need for fast decayssuch a particle
could be entirely stable. Another similar possibility is that has a
huge millicharge in the neighborhood of 105 103 e. Such a scheme
would require the violation of charge conservation as the possibility of
neutrino charges within a simple extension of the Standard Model discussed above always gives charges to two neutrino avors; for e or
the required value is not tolerable (Sect. 15.8).
For the issues of stellar evolution, the only conceivable consequence
of such large electromagnetic interaction cross sections would be
a reduced contribution to the energy transfer in SNe because of the
reduced mean free path. In the study discussed in Sect. 13.6 one should
have included the possibility of only two eective avors! Still, there
is little doubt that large cross sections could be accommodated in
what one knows about SNe today.
d) Astrophysical Bounds on Dipole and Transition Moments
For all neutrinos with a mass below a few keV a very restrictive limit
on dipole or transition magnetic or electric moments arises from the absence of anomalous neutrino emission from the cores of evolved globularcluster stars, notably of red-giant cores just before helium ignition
12
(Sect. 6.5.6). One nds a limit <
310 B which applies to Dirac
and Majorana neutrinos, and which does not allow for a destructive
interference between electric and magnetic amplitudes.
Neutrino transition moments would reveal themselves by radiative
decays. Because the decay rate involves a phase-space factor m3 this
method is suitable only for large masses. In the cosmologically allowed
range with m <
30 eV, the only radiative limit which can compete
with the globular-cluster bound is from the cosmic diuse background
radiations (Fig. 12.21). In fact, Sciama has proposed a scheme where a
28.9 eV neutrino with a radiative decay time corresponding to a tran-
579
Summary
Neutrinos with nonstandard interactions may well saturate the experimental mass limits, and may have a variety of novel properties. However, it is nearly impossible to derive generic constraints on quantities 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 sufciently 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 limits. If a concrete conjecture turns up, or a specic 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
Units and Dimensions
In the astrophysical context, frequently occurring units of length are
centimeters, (light) seconds, light years, and parsecs. Conversion factors are given in Tab. A.1. For example, 1 pc = 3.26 ly.
Using both centimeters and (light) seconds as units of length implies
a system of units where the speed of light c is dimensionless and equal to
unity. In this book I always use natural units where Plancks constant h
581
1G
1 erg/cm3
= 1.953102 eV2 = 0.502108 cm2 , (A.1)
4
where I have converted erg and cm1 into eV according to Tab. A.2.
The energy density of a magnetic eld of strength 1 G is, therefore,
1
(1.953102 eV2 )2 = 1.908104 eV4 = 3.979102 erg cm3 =
2
(1/8) erg cm3 . For a further discussion of electromagnetic units see
Jackson (1975).
It is sometimes useful to measure very strong magnetic elds in
terms of a critical eld strength Bcrit which is dened by the condition
that the quantum energy corresponding to the classical cyclotron frequency h
(eB/me c) of an electron equals its rest energy me c2 so that in
natural units
Bcrit = m2e /e.
(A.2)
1
2.9981010
1.3101011
1.5191015
1.4151024
0.9481027
0.8521048
cm1
0.3341010
1
4.369
0.507105
0.4721014
0.3161017
2.8431037
s1
cm1
K
eV
amu
erg
g
s1
1
2.9981010
0.9461018
3.081018
ly
1.061018
0.317107
1
3.26
s
0.3341010
1
3.156107
1.028108
0.3251018
0.973108
0.307
1
pc
0.7641011
0.2289
1
1.160104
1.0811013
0.7241016
0.6511037
K
0.6581015
1.973105
0.862104
1
0.931109
0.6241012
0.5611033
eV
0.7071024
2.1181014
0.9261013
1.074109
1
0.670103
0.6021024
amua
1.0551027
3.1611017
1.3811016
1.6021012
1.492103
1
0.8991021
erg
cm
s
ly
pc
cm
1.1731048
0.3521037
1.5371037
1.7831033
1.6611024
1.1131021
1
582
Appendix A
Appendix B
Neutrino Coupling Constants
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 eective four-fermion coupling
which is parametrized in terms of the Fermi constant and the weak
mixing angle
GF = 1.166105 GeV2 ,
sin2 W = 0.2325 0.0008.
(B.1)
2 GF = 2
= 2
,
(B.2)
2
2
mW sin W
mZ sin W cos2 W
where mZ = 91.2 GeV and mW = 80.2 GeV.
The eective charged-current interaction between nucleons and leptons is written in the form
GF
Hint = p (CV CA 5 )n (1 5 ) ,
2
(B.3)
where the j are the proton, neutron, charged-lepton, and the corresponding neutrino eld. The vector-current coupling constant is CV = 1
while the axial-vector coupling for free nucleons is CA = 1.26. However,
in large nuclei this value is suppressed somewhat, and the commonly
used value for nuclear matter is CA = 1.0 (e.g. Castle and Towner 1990).
This quantity would be relevant, for example, for reactions in supernova cores and neutron stars.
583
584
Appendix B
(B.4)
Table B.1. Neutral-current couplings for the eective Hamiltonian Eq. (B.4)
in vacuum.
Fermion f
Electron
Proton
Neutron
Neutrino (a )
Neutrino
CV
1
e
+ 2 + 2 sin2 W
,
21 + 2 sin2 W
e,,
+ 12 2 sin2 W
e,,
12
a
+1
b=a
+ 21
CA2
CA
CV2
+ 12
0.9312 0.25
12
0.0012 0.25
+1.37/2 0.0012 0.47
1.15/2 0.25 0.33
+1
1
1
1
+2
0.25 0.25
Appendix C
Numerical Neutrino
Energy-Loss Rates
In normal stars with densities below nuclear there are four main reactions that contribute to the energy loss by neutrino emission:
pl
e
e+ e
e (Z, A)
(Z, A) e
Plasma process,
Photoneutrino process,
Pair annihilation,
Bremsstrahlung.
(C.1)
C.1
Plasma Process
Widely used formulae for the plasma process are those of Beaudet, Petrosian, and Salpeter (1967), Munakata, Kohyama, and Itoh (1985), and
Schinder et al. (1987), which all agree with each other to better than 1%
if the same eective coupling constants are used. All of these rates are
8
poor approximations for T <
10 K which is relevant for low-mass stars
because they were optimized for higher temperatures. Itoh et al. (1989)
have attempted to improve the accuracy at low temperatures, and Blinnikov and Dunina-Barkovskaya (1994) gave rates which were optimized
for low-mass stars but fail for temperatures above about 108 K. At high
temperatures and densities, a poor approximation to the photon dispersion relation was used in all of these works (Braaten 1991) whence
none of these rates are satisfactory. A new t by Itoh et al. (1992) still
contains islands in the -T -plane with errors of several 10%.
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. e 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.
C.2
Beaudet, Petrosian, and Salpeter (1967) provided analytic approximations for the photoneutrino and pair-annihilation processes. Dicus
(1972) gave global correction factors to these rates to include neutralcurrent eects. Schinder et al. (1987) numerically recalculated the emission rates in the standard model and found good agreement with the
BPS formulae together with the Dicus correction factors. They supplemented the BPS rates for the temperature range 1010 1011 K.
An alternate set of approximation formulae was provided by Itoh
et al. (1989) who improved on their previous work (Munakata, Kohyama, and Itoh 1985). In Fig. C.2 I show the relative deviation between
the Itoh et al. (1989) with the Schinder et al. (1987) rates. The total
587
Fig. C.2. Deviation between the Schinder et al. (1987) and the Itoh et al.
(1989) rates for the photoneutrino and pair-annihilation processes. Compared are the total emission rates where the plasma rate of Haft, Raelt and
Weiss (1994) and the bremsstrahlung rate (helium) of Itoh and Kohyama
(1983) were used. The contours indicate were the individual processes dominate (Fig. C.1).
energy-loss rates for the photo and pair process was calculated according to these authors, while in each case the plasma rate of Haft, Raelt,
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 dierence between the
rates in this regime does not appear to be of much practical signicance.
Another analytic approximation formula for the pair process was
derived by Blinnikov and Rudzski (1989).
588
C.3
Appendix C
Bremsstrahlung
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 dierent processes dominate are for iron.
C.4
589
The total neutrino energy-loss rate for helium is shown in Figs. C.4C.6
where the photoneutrino and pair-annihilation rates are from Schinder
et al. (1987), the plasma process from Haft, Raelt, and Weiss (1994),
and bremsstrahlung (helium) from Itoh and Kohyama (1983)
Fig. C.4. Contour plot for the total neutrino energy-loss rate per unit
mass for helium. The thin contours are at intervals of a factor of 10 for .
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.
Appendix D
Characteristics of Stellar
Plasmas
D.1
Normal Matter
D.1.1
The material encountered in stars is usually in a state of thermal equilibrium. In the absence of strong magnetic elds, the plasma is entirely
characterized by its temperature T , mass density , and a set of chemical composition parameters X, Y , X12 , etc. which determine the mass
fractions of the elements 1 H, 4 He, 12 C, and so forth. The mass fraction
of all elements heavier than helium (metals) is denoted by Z.
The number density of a species with mass fraction Xj , atomic
weight Aj , and charge Zj e is given by
nj = (/mu ) Xj /Aj ,
(D.1)
Zj n j =
X j Zj
=
,
mu j Aj
e mu
(D.2)
where e is the mean molecular weight per electron, not to be confused with the electron chemical potential. (Strictly speaking ne =
ne ne+ , the number density of electrons minus that of positrons.)
99
The proton and neutron mass are 0.9383 and 0.9396 GeV, respectively. An exact
translation between mass and number density thus requires taking nuclear binding
energies into account whence the Aj are not exact integers. For the purposes of this
book these dierences are negligible.
591
592
Appendix D
(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 .
593
The nuclei in normal stellar matter are always nonrelativistic; relativistic corrections begin to be important only in neutron stars. The electrons, on the other hand, tend to be at least partially relativistic. Even
at the center of the Sun at a temperature of 1.3 keV, a typical thermal
electron velocity is about 9% of the speed of light. In Fig. D.2 contours
for the thermal average v 2 1/2 are shown in the -T -plane. The loci of
the stellar models of Fig. D.1 are also indicated. For low-mass stars, the
electrons are mildly relativistic, although a nonrelativistic treatment is
often enough as a rst approximation.
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.
D.1.3
Electron Degeneracy
The phase-space occupation numbers of fermions in thermal equilibrium are characterized by a Fermi-Dirac distribution
fp =
1
e(Ep )/T
+1
(D.4)
594
Appendix D
nf =
2 d3 p
(2)3
1
e(Ep )/T + 1
1
e(Ep +)/T + 1
(D.5)
where the second term represents antifermions and the factor 2 is for the
two spin degrees of freedom. Again, the fermion density is understood
to mean the density of fermions minus that of antifermions.
At vanishing temperature, fp becomes a step function ( Ep ).
If > 0 so that nf > 0, i.e. an excess of fermions over antifermions,
there are no antifermions at all at T = 0. The fermion integral yields
nf = p3F /3 2 ,
(D.6)
(D.7)
(D.8)
1
e(Ekin )/T
+1
(D.9)
595
Fig. D.3. Contours for the electron chemical potential . The solid lines are
marked with the relevant value for , the dotted lines with me .
596
Appendix D
/T = ( m)/T
(D.10)
Plasma Frequency
(D.11)
(Ye )1/2
,
[1 + (1.019106 Ye )2/3 ]1/4
(D.12)
Screening Scale
(D.13)
4 2
4 Xj Zj2
.
=
Z nj =
T j j
T mu j
Aj
(D.14)
597
2
For only one species of ions with charge Ze one has ki2 = Z kD
. Numerically,
(D.15)
where is in units of g cm3 and T8 = T /108 K. Contours in the -T plane are shown in Fig. D.5.
(D.16)
(D.17)
598
Appendix D
with the mass density in units of g cm3 and T8 = T /108 K. Recall that
for a single nuclear species Ye = Z/A (atomic weight A).
The plasma is weakly coupled for <
1, it is in the liquid metal
<
phase for 1 < 178, and forms a body-centered cubic lattice for
> 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. Contours for in the -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 . 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 astrophysical sites that are important in the main body of the book are summarized in Tab. D.1.
degenerate
weakly coupled
108 K
= 8.6 keV
106 g cm3
4
He
3.01029 cm3
409 keV
144 keV
18 keV
0.57
ions
kS = ki
= 222 keV
nondegenerate
nonrelativistic
108 K
= 8.6 keV
104 g cm3
4
He, 12 C, 16 O
3.01027 cm3
88 keV
7.6 keV
2.0 keV
0.12
electrons + ions
2
+ ki2 )1/2
kS = (kD
= 27 keV
Core of
HB stars
Red-giant core
just before
helium ignition
nondegenerate
nonrelativistic
Temperature
1.55107 K
= 1.3 keV
Density
156 g cm3
Composition
X = 0.35
Electron density
6.31025 cm3
Fermi momentum 24.3 keV
Fermi energyb
0.58 keV
Plasma frequency 0.3 keV
Plasma coupling
= 0.07
Debye screening
electrons + ions
2
+ ki2 )1/2
kS = (kD
= 9.1 keV
Characteristic
Center of
standard
solar model
degenerate
strongly coupled
3106 2107 K
= 0.31.7 keV
1.8106 g cm3 (a )
12
C, 16 O
5.31029 cm3
495 keV
200 keV
23 keV
14422
(strong
screening)
White dwarf
600
Appendix D
D.2
Nuclear Matter
D.2.1
For the topics discussed in this book, the properties of hot nuclear
matter in a young supernova core are of great interest. The relevant
range of densities and temperatures is about 31012 31015 g cm3
and 3100 MeV, respectively. The properties of matter at such conditions is determined by its equation of state which takes the nuclear
interaction fully into account. However, in order to gain a rough understanding of the behavior of the main constituents of the medium
(protons, neutrons, electrons, and electron neutrinos) it is worthwhile
to study a simple toy model where these particles are treated as ideal
Fermi gases.
To this end, neutrinos and electrons are treated as massless. Their
dispersion relation is dominated by the interaction with the medium.
Because they interact only by electroweak forces their eective mass
is always much smaller than their energies.
D.2.2
The reaction e p n e which establishes equilibrium is fast compared to other relevant time scales. Therefore, the relative abundances
of n, p, e, and e are determined by the conditions of kinetic and chemical equilibrium. The physical condition of the medium is then determined by the baryon density nB , the temperature T , and the condition
of electric charge neutrality
nB = nn + np
np = ne
e + p = n + e
Baryon density,
Charge neutrality,
equilibrium.
(D.18)
Lepton conservation.
(D.19)
601
(D.20)
(D.21)
(D.22)
(D.23)
x4n
.
4 (1 + x2n )
(D.24)
This result may be expressed in terms of the usual composition parameters 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
(
Yp =
xB
2
)3
(1 Yp )2
.
[1 + x2B (1 Yp )2/3 ]3/2
(D.25)
602
D.2.4
Appendix D
Hot Nuclear Matter
In a supernova core right after collapse the temperature is so high (several tens of MeV) that the nucleons are nearly nondegenerate. Moreover, the neutrinos are trapped so that locally a xed value for YL
determined by initial conditions is assumed. Most of the lepton number will reside in electrons, causing the proton concentration Yp to be
approximately equal to YL . A more accurate determination requires a
numerical solution of Eqs. (D.18) and (D.19).
In Figs. D.7 (a)(c) the results of such an exercise are presented
for YL = 0.3, which is a typical value for the material in a SN core just
after collapse. The proton concentration Yp , the dierence between the
neutron and proton chemical potentials, and the degeneracy parameters
for neutrons and protons are shown. In each case, a solid line refers
to the assumption of an eective nucleon mass as in Fig. 4.10 while
a dotted line refers to the vacuum mass. As expected, the reduced
eective mass increases somewhat the mediums degeneracy.
Because n p = e e , the contours of Fig. D.7 (b) also give
the dierence between the surfaces of the electron and neutrino Fermi
seas. Note that the leptons are much more degenerate than the nucleons
because they are essentially massless. This remark does not apply to the
upper-left corner of the plots where actually an excess of antineutrinos
is enforcedthere are more protons than leptons (Yp > YL )!
Fig. D.7. (a) Contours for Yp in hot neutron-star matter with YL = 0.3.
Solid lines for the eective nucleon mass as in Fig. D.7, dotted lines for the
vacuum mass.
603
604
Appendix D
Fig. D.9. Ratio of the suppression factor of Fig. D.8 between the case with
an eective nucleon mass and with the vacuum one.
(D.26)
605
YL = 0.3 and the eective nucleon mass of Fig. 4.10. In Fig. D.9 the
ratio of this factor between the case of an eective nucleon mass and
the vacuum one are shown, i.e. the additional Pauli suppression of the
neutrino scattering rate from using an eective nucleon mass.
References
Prexes 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.
Abbott, L. F., de R
ujula, A., and Walker, T. P. 1988, Nucl. Phys. B,
299, 734.
Abbott, L. F., and Sikivie, P. 1983, Phys. Lett. B, 120, 133.
Abdurashitov, J. N., et al. 1994, Phys. Lett. B, 328, 234.
Accetta, F. S., Krauss, L. M., and Romanelli, P. 1990, Phys. Lett. B,
248, 146.
Accetta, F. S., and Steinhardt, P. J. 1991, Phys. Rev. Lett., 67, 298.
Achkar, B., et al. 1995, Nucl. Phys. B, 434, 503.
Acker, A., and Pakvasa, S. 1994, Phys. Lett. B, 320, 320.
Acker, A., Pakvasa, S., and Raghavan, R. S. 1990, Phys. Lett. B, 238,
117.
Adams, E. N. 1988, Phys. Rev. D, 37, 2047.
Adams, J. B., Ruderman, M. A., and Woo, C.-H. 1963, Phys. Rev., 129,
1383.
Adler, S. L. 1971, Ann. Phys. (N.Y.), 67, 599.
Aharonov, Y., Avignone III, F. T., and Nussinov, S. 1988a, Phys. Lett.
B, 200, 122.
Aharonov, Y., Avignone III, F. T., and Nussinov, S. 1988b, Phys. Rev.
D, 37, 1360.
Aharonov, Y., Avignone III, F. T., and Nussinov, S. 1989, Phys. Rev.
D, 39, 985.
Ahrens, L. A., et al. 1985, Phys. Rev. D, 31, 2732.
Ahrens, L. A., et al. 1990, Phys. Rev. D, 41, 3301.
Akhmedov, E. Kh. 1988a, Yad. Fiz., 48, 599 (Sov. J. Nucl. Phys., 48,
382).
Akhmedov, E. Kh. 1988b, Phys. Lett. B, 213, 64.
Akhmedov, E. Kh., and Berezin, V. V. 1992, Z. Phys. C, 54, 661.
606
References
607
608
References
Anselm, A. A. 1982, Pisma 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
References
611
612
References
References
613
614
References
References
615
616
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 H
uckel, E. 1923, Phys. Z., 24, 185.
DeglInnocenti, S., Fiorentini, G., and Lissia, M. 1995, Nucl. Phys. B
(Proc. Suppl.), 43, 66.
DeglInnocenti, 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 Raelt, G. G. 1995, Phys. Rev. D, 52, 2581.
DOlivo, J. C., Nieves, J. F., and Pal, P. B. 1989, Phys. Rev. D, 40,
3679.
DOlivo, 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, Pisma Astron. 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
References
619
620
References
References
621
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 Masso, E. 1986, Phys. Lett. B, 173, 237.
Grifols, J. A., and Masso, E. 1989, Phys. Rev. D, 40, 3819.
Grifols, J. A., and Masso, E. 1990a, Phys. Lett. B, 242, 77.
Grifols, J. A., and Masso, E. 1990b, Nucl. Phys. B, 331, 244.
Grifols, J. A., Masso, E., and Peris, S. 1988a, Phys. Lett. B, 207, 493.
Grifols, J. A., Masso, E., and Peris, S. 1988b, Phys. Lett. B, 215, 593.
Grifols, J. A., Masso, E., and Peris, S. 1989a, Phys. Lett. B, 220, 591.
Grifols, J. A., Masso, E., and Peris, S. 1989b, Mod. Phys. Lett. A, 4,
311.
Grifols, J. A., Masso, E., and Peris, S. 1994, Astropart. Phys., 2, 161.
Grifols, J. A., Masso, 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, Masters Thesis, University of Munich, unpublished.
Haft, M., Raelt, 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
References
623
624
References
References
625
626
References
Kolb, E. W., and Turner, M. S. 1990, The Early Universe (AddisonWesley, Reading, Mass.).
Kolb, E. W., et al. 1991, Phys. Rev. Lett., 67, 533.
Kopysov, Yu. S., and Kuzmin, V. A. 1968, Can. J. Phys., 46, S488.
Kohyama, Y., Itoh, N., and Munakata, H. 1986, Ap. J., 310, 815.
Kohyama, Y., et al. 1993, Ap. J., 415, 267.
Konoplich, R. V., and Khlopov, M. Yu. 1988, Yad. Fiz., 47, 891 (Sov.
J. Nucl. Phys., 47, 565).
Koshiba, M. 1992, Phys. Rep., 220, 229.
Kosteleck
y, V. A., Pantaleone, J., and Samuel, S. 1993, Phys. Lett. B,
315, 46.
Kovetz, A., and Shaviv, G. 1994, Ap. J., 426, 787.
Krakauer, D. A., et al. 1990, Phys. Lett. B, 252, 177.
Krakauer, D. A., et al. 1991, Phys. Rev. D, 44, R6.
Krastev, P. I. 1993, Phys. Lett. B, 303, 75.
Krastev, P. I., and Petcov, S. T. 1994, Phys. Rev. Lett., 72, 1960.
Krastev, P. I., and Smirnov, A. Yu. 1994, Phys. Lett. B, 338, 282.
Krauss, L. M. 1984, Phys. Rev. Lett., 53, 1976.
Krauss, L. M. 1991, Phys. Lett. B, 263, 441.
Krauss, L. M., Moody, J. E., and Wilczek, F. 1984, Phys. Lett. B, 144,
391.
Krauss, L. M., and Tremaine, S. 1988, Phys. Rev. Lett., 60, 176.
Krauss, L. M., and Wilczek, F. 1986, Phys. Lett. B, 173, 189.
Krauss, L. M., et al. 1985, Phys. Rev. Lett., 55, 1797.
Krauss, L. M., et al. 1992, Nucl. Phys. B, 380, 507.
Kuhn, J. R. 1988, in: J. Christensen-Dalsgaard and S. Frandsen (eds.),
Advances in Helio- and Asteroseismology (Reidel, Dordrecht).
Kunihiro, T., et al. 1993, Prog. Theor. Phys. Suppl., No. 112.
Kuo, T. K., and Pantaleone, J. 1988, Phys. Rev. D, 37, 298.
Kuo, T. K., and Pantaleone, J. 1989, Rev. Mod. Phys., 61, 937.
Kuo, T. K., and Pantaleone, J. 1990, Phys. Lett. B, 246, 144.
Kuznetsov, A. V., and Mikheev, N. V. 1993, Phys. Lett. B, 299, 367.
Kwong, W., and Rosen, S. P. 1994, Phys. Rev. Lett., 73, 369.
Kyuldjiev, A. V. 1984, Nucl. Phys. B, 243.
Lagage, P. O., et al. 1987, Phys. Lett. B, 193, 127.
Lam, W. P., and Ng, K.-W. 1991, Phys. Rev. D, 44, 3345.
Lam, W. P., and Ng, K.-W. 1992, Phys. Lett. B, 284, 331.
Lamb, D. Q., and van Horn, H. M. 1975, Ap. J., 200 306.
Landau, L. D. 1946, Sov. Phys. JETP, 16, 574.
Landau, L. D., and Lifshitz, E. M. 1958, Statistical Physics (AddisonWesley, Reading, Mass.).
References
627
628
References
References
629
630
References
References
631
Ogelman,
H., and Finley, J. 1993, Ap. J., 413, L31.
Ogelman, H., Finley, J., and Zimmermann, H. 1993, Nature, 361, 136.
Okun, L. B. 1969, Yad. Fiz., 10, 358 (Sov. J. Nucl. Phys., 10, 206).
Okun, L. B. 1986, Yad. Fiz., 44, 847 (Sov. J. Nucl. Phys., 44, 546).
Okun, L. B. 1988, Yad. Fiz., 48, 1519 (Sov. J. Nucl. Phys., 48, 967).
Olive, K. A., Schramm, D., and Steigman, G. 1981, Nucl. Phys. B, 180,
497.
Olive, K. A., et al. 1990, Phys. Lett. B, 236, 454.
Oraevski, V. N., and Semikoz, V. B. 1984, Zh. Eksp. Teor. Fiz., 86,
796 (Sov. Phys. JETP, 59, 465).
Oraevski, V. N., and Semikoz, V. B. 1985, Yad. Fiz., 42, 702 (Sov. J.
Nucl. Phys., 42, 446).
Oraevski, V. N., and Semikoz, V. B. 1987, Physica, 142A, 135.
Oraevski, V. N., and Semikoz, V. B. 1991, Phys. Lett. B, 263, 455.
Oraevski, V. N., and Semikoz, V. B., and Smorodinski, Ya. A. 1986,
Pisma Zh. Eksp. Teor. Fiz., 43, 549 (JETP Lett., 43, 709).
Otten, E. W. 1995, Neutrino 94, Nucl. Phys. B (Proc. Suppl.), 38, 26.
Overduin, J. M., and Wesson, P. S. 1993, Ap. J., 414, 449.
Overduin, J. M., Wesson, P. S., and Bowyer, S. 1993, Ap. J., 404, 460.
Page, D., and Applegate, J. H. 1992, Ap. J., 394, L17.
Pakvasa, S., Simmons, W. A., and Weiler, T. J. 1989, Phys. Rev. D,
39, 1761.
Pal, P. B., and Pham, T. N. 1989, Phys. Rev. D, 40, 259.
Pal, P. B., and Wolfenstein, L. 1982, Phys. Rev. D, 25, 766.
632
References
References
633
634
References
References
635
636
References
References
637
638
References
References
639
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.), Observational 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-Chi`eze, S., and Lopes, I. 1993, Ap. J., 408, 347.
Turck-Chi`eze, 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. Astrophys., 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, IAUSymposium No. 42 (Reidel, Dordrecht).
Vassiliadis, G., et al. 1995, Proc. Int. Symp. Strangeness and Quark
Matter , Sept. 15, 1994, Crete, Greece (World Scientic, Singapore).
640
References
References
641
Acronyms
The acronyms listed below do not include names of experiments such
as GALLEX or IMB, and also do not include names of laboratories
such as CERN or SLAC. These acronyms really play the role of proper
names; usually nobody quite remembers what they stand for.
AGB
AU
BBN
BC
BP
BS
CC
CKM
CL
CM
CMB
CMBR
CNO
CO
D
DAV
DFSZ
EOS
FTD
GIM
GUT
HB
IAU
IMF
KII
pc
KSVZ
Acronyms
L
l.h.
l.h.s.
ly
LMC
LSP
mfp
MS
MSW
NC
ND
T
TL
TO
OPE
PQ
QED
QCD
RG
RGB
r.h.
r.h.s.
SGB
SN
SNe
SNU
SNu
SNBO
UT
VVO
WD
643
longitudinal
left-handed
left-hand side (of an equation)
light year
Large Magellanic Cloud
lightest supersymmetric particle
mean free path
main sequence
Mikheyev-Smirnov-Wolfenstein
neutral current
nondegenerate
transverse
Turck-Chi`eze and Lopes
main-sequence turno
one pion exchange
Peccei-Quinn
quantum electrodynamics
quantum chromodynamics
red giant
red-giant branch
right-handed
right-hand side (of an equation)
sub-giant branch
supernova
supernovae
solar neutrino unit
supernova unit
Supernova Burst Observatory
universal time
Voloshyn-Vysotski-Okun
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 (, , , , etc.),
three-indices and avor indices in Latin letters (i, j, k, etc.).
Functions, Operators
Besides the usual functions and operators, the following convention may
be noteworthy.
x
ln
log
partial derivative /x
natural logarithm
logarithm base 10
Latin Symbols
a
ai
A
B
B
c
cj
cp
CV,A
Cj
d
d, D
Symbols
dp
e
ej
E
EF
f
fPQ
fa
f
fp
F
F
F
g
ga
g10
G
GN
GF
H
H0
H
k
kB
kD
ki
kS
kTF
l
L
L
L
m
ma
me
m
645
646
Symbols
mN
nucleon mass (938 MeV)
mu
atomic mass unit (931 MeV)
mPl Planck mass (1.2211019 GeV)
mW W -boson mass (80.2 GeV)
mZ
Z-boson mass (91.2 GeV)
M
absolute magnitude of a star, mass matrix of quarks or neutrinos
Mbol absolute bolometric magnitude of a star
M
mass of a star, matrix element
M solar mass (1.991033 g)
n
particle density
nB
baryon density
ne,p,n, electron, proton, neutron, neutrino density
nrefr index of refraction
N
nucleon, number of degenerate vacua in QCD
N matrix element of neutrino current
p
proton, momentum of a particle, pressure, dimensionless power
index
pF
Fermi momentum
P
four-momentum of a particle, stellar pulsation period
q
momentum transfer, charge of a particle
Q
energy loss or generation rate per unit volume
r
radial coordinate
R
radius of a star
R
solar radius (6.961010 cm)
s
entropy density, dimensionless power index, square of CM energy, dimensionless spin-structure function
S
static or dynamical structure function
t
time
T
temperature
T7
T /107 K
T8
T /108 K
T30
T /30 MeV
U
total internal energy of a star, four-velocity
v
velocity, vacuum expectation value of a Higgs-like eld
V
potential energy, volume
w
up/strange quark mass ratio
W
W -boson
WK,K Transition probability from K to K .
x
dimensionless energy (/T ), spatial coordinate, dimensionless
Fermi momentum (pF /m)
X
mass fraction of hydrogen
Symbols
Xj
Y
Yj
Ye
YL
z
Z
Z, Z
647
Greek Symbols
a,
s
ij
ij
648
B
N
e
0
14
15
P
0
Symbols
Bohr magneton (e/2me )
nuclear magneton (e/2mp )
electron mean molecular weight (e Ye1 ), electron chemical
potential
neutrino, Lorentz index, dimensionless power index
dimensionless correction factor, dimensionless axion-photon coupling constant
= 3.1415 . . ., pion
mass density, density matrix
density matrix for antineutrinos
nuclear density (31014 g cm3 )
/1014 g cm3
/1015 g cm3
scattering cross section, Pauli matrix, spin operator
lifetime of a particle, duration of a stellar evolution phase, Pauli
matrix, optical depth
Higgs eld
vacuum expectation value of Higgs eld
majoron, majoron eld
fermion eld
column vector of fermion elds (several avors)
energy of a particle, energy transfer in reactions
plasma frequency
zero-temperature plasma frequency
cosmic density parameter, matrix of energies for mixed neutrinos
Subject Index
A
-Ori 189
A665, A1413, A2218, A2256 (galaxy
clusters) 49092
adiabatic index 396
adiabatic oscillations
neutrino oscillations
adiabatic temperature gradient 1012
adiabaticity parameter 299
AGB (asymptotic giant branch) 27, 35,
61
aligned rotator 185
AMANDA xvi, 558
anapole moment 26870
Andromeda 446
angular anomaly of SN 1987A neutrino
signal 422, 42830
antimatter supernova 497
arions 18790
automatic symmetry 533
asymptotic giant branch 27, 35, 61
atmospheric neutrinos 2903
atomic mass unit 582, 591
axio-electric eect 100
axion
pseudoscalar boson
bounds 191f, 53941
cavity experiment 177f, 191f, 544
cosmic density 491f, 528, 54144
cosmic thermal production 541
couplings
electron 536f
fermions general 5302
gluon 159, 527, 530
nucleon 53638
photon 167f, 176, 191f, 536
decay 167f, 191f
decay constant 526f, 529, 535
direct search
decay photons of cosmic axions
491f
galactic (cavity search) 177f, 191f,
544
laboratory 182, 184f, 191f
solar 100, 181, 191f, 462
SN 1987A 483
mass 527, 535
mixing with pion 527f, 534
models
DFSZ 53540
KSVZ 52830, 54244
other 53436
Nambu-Goldstone boson 190f, 52830
quantum gravity 532f
solution to strong CP problem 527f
sphere 507
axio-recombination 100
B
decay 255
decay 257f, 560f
Baade-Wesselink method 73
Baksan scintillator telescope (BST)
414f, 417, 420
band-structure eects 58, 106f
baryonic force 114
beam-stop neutrinos 456f
Betelgeuse 189
big-bang nucleosynthesis: bounds
majorons 259f, 561f
millicharged particles 522f, 566f
neutrino
Dirac dipole moment 277f
mass 259f
time variation
Fermis constant 546
Newtons constant 547f
649
650
binary pulsar PSR 1913+16
birefringence
neutrino avor 195, 281
photon in media 180f, 194f, 208
vacuum 1835, 187f, 190f, 554
Bloch state 155
blue loop 32, 3941
bolometric correction 62
bolometric magnitude: denition 28
Boltzmanns constant 580
Boltzmann collision equation (for mixed
neutrinos) 32331
BOREXINO 343, 371, 393f
bounce and shock 399, 4015
bound-free transition 100f
branching ratio 449
Brans-Dicke theory 548
bremsstrahlung
classical limit 14347
electronic 101-9
nucleonic
axial-vector vs. vector 127, 142
bubble phase 159
meson condensate 15557
neutrino pairs 12730
pseudoscalars 11926, 14345, 148f
suppression at high density 148f
quark matter 157f
brown dwarfs 9, 24
bubble phase 159
bump on RGB 61, 64f
C
Cabbibo mixing 261, 285
Cabbibo-Kobayashi-Maskawa matrix
262
carbon-burning stars 2
cavity axion experiment 177f, 191f, 544
Cepheids 40f
Chandrasekhar limit 9, 25, 37, 42, 396
charged current
collision term (in neutrino
oscillations) 331
coupling constants 583
Hamiltonian 329, 583
reactions 1, 58f, 15254
charge quantization 56467
chemical potential 59396, 600f
Subject Index
Cherenkov
eect 194, 216, 238f
detectors IMB, Kamiokande,
Superkamiokande
chirality 160, 253
chiral symmetry 166, 529
chlorine detector (solar neutrinos) 342f,
35762, 371
CHORUS experiment 289f
CKM matrix 262
clump giants 76, 80
CM energy 92
CNO cycle 346f
COBE satellite 191
coherent neutrino scattering 396, 446
collision equation for mixed neutrinos
32331
color-magnitude diagram
evolutionary track 31
globular cluster M3 27
globular-cluster observables 61
populations of stars 41
zero-age main sequence 26
Compton process 20f, 9198, 108
conduction 11, 22, 78
convection
description 12f
helium dredge-up 66, 77
mixing length 12, 16, 351, 549f
RGB bump 13, 61, 64f
superadiabatic 12
supernova 4025, 444
core
red giant red giant core mass
supernova supernova core
correlations: Coulomb plasma 22225
cosmic background uxes
-rays 487
microwaves 191, 488, 546, 562
neutrinos
core-collapse SNe 447, 487f
hydrogen-burning stars 487
primordial neutrinos: cosmic
background
photons 489f
cosmic strings 191, 54244
cosmic structure formation 258f
cosmion xvi
Cotton-Mouton eect 180, 183
Subject Index
Coulomb gauge 204
Coulomb propagator in plasma 221f,
226
CP symmetry 241, 246, 262, 271, 453,
52426
Crab Nebula 38, 56f
critical eld strength 581
crossing relation 323
current algebra 167, 535
D
Cepheids 40f
Scuti stars 41
damping neutrino oscillations:
damping by collisions
DA Variables 53
dark matter xvxvi, xviii, 22, 24, 177f,
258f, 528, 54244, 558
decoherence 311, 313f
degeneracy parameter 124, 595f, 603
delayed explosion mechanism 399,
4025, 426, 436f, 522
Delbr
uck scattering 183
deleptonization burst
prompt neutrino burst
deleptonization: supernova core 339,
398, 401, 408
density matrix (neutrino avor)
collision equation 319, 32331
denition 284, 31516
derivative coupling 119, 531f, 536, 563
detailed balance 129, 137, 146
DFSZ axions 53540
dielectric permittivity 173, 208f
diusion
elements gravitational settling
neutrinos neutrino opacity
dipole moment
neutrino dipole moments
neutron electric 52426
other particles 525
units 525, 581
Dirac matrices 270
Dirac neutrinos neutrino mass,
neutrinos: right-handed
direct explosion mechanism 399, 401f
651
dispersion
neutrino dispersion, photon
dispersion
general theory 193202
double beta decay decay
DUMAND xvi, 558
dynamical structure function
structure function
E
Einstein x-ray satellite 5658, 189
electron charge 581
electro-Primako eect 172f
energy conservation in stars 10f
energy-loss argument
analytic treatment 1416
applied to
HB stars 7982, 98f, 1079, 112,
176, 236
neutron stars 59f
red giants 8387, 1079, 236
red supergiants 2
Sun 16f, 20f, 109f, 175f
supernova core 50123
white dwarfs 4752
ZZ Ceti stars 5254
introduction 35, 21f
numerical treatment 16f, 4752, 81,
8387, 108f, 50811, 51316
energy-loss rate
bremsstrahlung
electronic 1025
nucleonic 1216, 12830, 14749,
156f
Compton process 9698
plasma process 23236
Primako process 169, 174
quark matter 157f
URCA processes 153f
energy-momentum conservation 121,
200
energy transfer
convection, opacity
equation 1113
particle bounds
general argument 4f, 17f
Sun 20f
SN 1987A 503, 50611
652
energy transfer (contd)
radiative 4, 1719, 131
WIMPs 22
equation of state
impact on SN neutrino signal 412f
normal stars 6
nuclear 600
equivalence principle: tests 113, 498,
554f
escape velocity
galaxy 26
globular clusters 25
Euler-Heisenberg Lagrangian 183
EXOSAT 56f
explosion mechanism
supernova: type II
F
Faraday eect 180
Fermi
constant 545f, 583
energy 594
momentum 594
Fermi-Dirac distribution 59396
forward scattering 19698
fth force 11216, 500
ne-structure constant
axion-nucleon 119
electromagnetic 580
new bosons 92
pion-nucleon 119
avor-changing neutral current 263f,
303, 326
avor conversion
neutrino oscillations
uence: denition 463
foe 402
form factor neutrino form factor
free-bound transition 100f
Frejus experiment 291f
FTD eld theory xviiixix, 105, 174,
241
fundamental length scale 500
G
G117B15A 53f
Subject Index
GALLEX 343f, 35759, 36264, 371,
373
gallium detectors 343f, 35759, 36264,
371
gauge
bosons: new 11416
invariance 2046
General Relativity gravitation
giant star
red giant
denition 10
GIM suppression 266f
globular cluster
red giants, HB stars, RR Lyrae
stars
ages 62f, 66, 55254
general properties 25f
M3 25, 27
observables
interpretation 7479
in the color-magnitude diagram
6065
observational results 7074
theoretical relations 6570
gluons 159, 525, 527, 530
g-mode helioseismology
gravitation
Brans-Dicke theory 548
equivalence principle: tests 113, 498f,
554f
graviton 116, 166, 191, 254
induced neutrino mixing 555
nonmetric theory 498
nonsymmetric extension 554
quantum 532f
Shapiro time delay 498, 554
time-varying GN 54654
gravitational settling
description 13f
globular-cluster ages 13, 66
red-giant envelope helium abundance
66, 74, 77
solar neutrino ux 14, 352, 354
graviton 116, 166, 191, 254
group velocity 199
GRS instrument 465f
Subject Index
653
I, J
IMB detector
atmospheric neutrinos 291f
SN 1987A neutrinos 41523
ination 542
ions
contribution to screening 220f, 596,
599
correlations 106, 22225
ion-sphere radius 223f, 597
isochrone 61f
Jupiter magnetic eld 556
K
mechanism 40
Kamiokande (Cherenkov detector)
atmospheric neutrinos 291f
dark-matter search xvi
neutrino detection
eciency 417
electron scattering 36567, 416f
proton absorption 416f
proton decay xvii
SN 1987A neutrinos 41823
solar neutrinos 343f, 36873, 383
Kelvin-Helmholtz cooling
Sun 8
supernova core 397, 400, 40713,
50123
Kibble mechanism 542
kick velocity (neutron stars) 44345
Klein-Gordon equation for mixed
neutrinos 28284, 294
Kramers-Kronig relation 198
KSVZ axions 52830, 54244
L
Lamb shift 565f
Landau damping 199, 214, 216f
Landau-Pomeranchuk-Migdal eect
145, 148f
Landau-Zener approximation 300
Langmuir wave
plasmon: longitudinal
Large Magellanic Cloud 39, 74, 414, 446
large-numbers hypothesis 545, 549
654
left-right symmetric model 268
leptonic force 11416, 500
lepton
fraction in nuclear matter 600, 6025
masses 252, 265
prole in SN core 398
level crossing 295f, 298
linear-response theory 2046, 208f
long-range force fifth force,
scalar particles, photon
long-wavelength limit 121, 128f, 14143
Lorentz
addition of velocities 500f
gauge 204
Lorentzian model for structure function
135, 14447
LPM eect 145, 148f
M
M1 transition 94, 102
M3 (globular cluster) 25, 27
magnetic dipole moment
neutrino dipole moments
units 581
magnetic eld
Earth 556
galactic 18890, 484, 499
large-scale cosmic 191
Jupiter 556
neutron stars 18588, 444f
primordial 278
Small Magellanic Cloud 556
solar 186, 37376, 387f, 554, 565
supernova 522
twisting 308f
white dwarfs 185f
magnetic oscillations
neutrino spin precession
magnetic permeability 208f
magnetogram 375
main sequence
helium 35, 592
internal stellar conditions 592, 599
lifetime 27, 552
location in color-magnitude diagram
27, 41, 61
turno 27, 6163
zero-age 26
Subject Index
Majorana neutrinos neutrino mass
majorons
pseudoscalar bosons
decay 560f
decay in media 195, 24850, 259f
interaction with supernova neutrinos
495
motivation 55862
neutrino decay: limits 56264
nucleosynthesis constraints 259f
solar neutrino decay 389
supernova physics 56264
mass fraction (of elements) 591f
mass loss 25, 34, 36
mass-radius relationship 9, 42
Maxwells equations 2024
mean molecular weight 591
meson condensate 59, 15557, 243
metallicity
-enhanced 78f
globular cluster 26, 34, 60, 66f, 78f
impact on opacity 12
Sun 24
Mexican hat 529, 542
Mikheyev-Smirnov-Wolfenstein eect
neutrino oscillations: resonant
millicharged particles 227f, 23236,
56467
misalignment mechanism 541f
mixing angle in medium 2947
mixing between particles
neutrino mixing, neutrino
oscillations
axion-photon 17981
axion-pion 527f, 534
general theory 26063
graviton-photon 191
gravitationally induced 555
mixing length 12, 16, 351, 549f
molecular chaos 319
monopoles xv
MSW
bathtub 3013
eect neutrino oscillations
solution of solar neutrino problem
38486
triangle 3013, 386, 434
Subject Index
multiple scattering
Cherenkov detector 366
nuclear medium 135, 14351, 511f
MUNU experiment 276
muon decay 263
N
Nambu-Goldstone bosons 165f, 250,
52931, 559f
NESTOR xvi, 558
neutral current
collision term in neutrino oscillations
323f
coupling constants 584
avor-changing 263f, 303, 326
Hamiltonian 95, 119, 136, 320, 531,
584
neutrality of matter 565
neutrino absorption neutrino opacity
neutrino charge: electric 227f, 23236,
240, 499, 522f, 56467
neutrino cooling
supernova core
red-giant core 8385
white dwarfs 4750
neutrino coupling constants: standard
583f
neutrino cross section
electrons 365f, 416
protons and nuclei 416
neutrino dark matter 25860
neutrino decay
neutrino radiative decay
fast invisible 24850, 389, 485, 559
standard-model 26367
neutrino dipole moments
astrophysical impact and bounds
early universe 277f
HB stars 82, 236
neutron stars 59f
red-giant core mass 86f, 236
supernova 277, 47779, 483, 521f
white dwarfs 4850
electric vs. magnetic 265f, 26871,
27375, 525
experimental bounds
neutrino radiative decay
scattering experiments 267, 275f
655
neutrino dipole moments (contd)
interaction structure 228, 265, 26871
medium induced 240
processes
plasmon decay 229, 23236, 274
scattering 272f
summary 27275
radiative decay
neutrino radiative decay
summary of limits 27579
theoretical prediction
Dirac vs. Majorana 265, 271, 451
left-right symmetric model 268
standard-model 265f
structure of interaction 265, 26871
neutrino dispersion
Cherenkov eect 238f
deection 247f
density-matrix treatment 321f
general theory 24147
inhomogeneous medium 442f
neutrino background 321f, 432f,
44042
nonisotropic medium 321, 432f,
44042
o-diagonal refractive index 322
relation for massive neutrinos in
media 295f
spin-ip scattering 162
neutrino emission: standard
solar neutrino flux, supernova core
astrophysical impact
red-giant core mass 83f
red supergiant 2
white-dwarf cooling 4750
historical perspective 1f
numerical rates 58590
specic processes
bremsstrahlung (electronic) 105-7,
588
bremsstrahlung (nucleonic) 12730,
157, 159
Compton (photoneutrino) 95f, 98,
586f
free-bound transition 100f
pair annihilation 99f, 586f
plasma 585f
656
neutrino form factor: electromagnetic
neutrino dipole moment
anapole moment 26870
charge 227f, 23236, 240, 268, 56467
charge radius 269f, 523
general theory 26772
in plasma 230f, 23741
neutrino lasing 389
neutrino lightcurve supernova core
neutrino mass
bounds
decay 257f
cosmological 25860
experimental 252, 25458
future supernovae 447f, 497
SN 1987A signal dispersion 426f
SN 1987A signal duration 51619
Dirac 253f, 51619
eective neutrino dispersion
inverted hierarchy 434
Majorana 253f, 25759
neutrino mixing
neutrino oscillations
bounds
atmospheric neutrinos 291
decay experiments 456f, 461
oscillation experiments 289f
r-process nucleosynthesis 43742
SN 1987A -rays 481
SN 1987A prompt burst 43234
SN 1987A signal duration 43436
supernova core: avor conversion
338
decay rates and dipole moments
26467
induced by
avor-changing neutral currents
303
gravity 555
mass matrix 262, 282
neutrino opacity 13235, 14951, 329f,
334f, 408, 51416
neutrino oscillations
damping by collisions
kinetic equation 32729, 33234
simple picture 31012
spin relaxation 313
time scale in SN core 33538
equations of motion 28284
Subject Index
neutrino oscillations (contd)
experimental searches 289f
historical introduction and overview
28082
in media
adiabatic limit 29799
analytic results 299300
homogeneous 296f
mixing angle 2947
neutrino background in supernovae
432f, 44042
oscillation length 285f, 289, 29497
primordial 322
resonant
analytic results 299f
description 281, 298
Landau-Zener approximation 300
level crossing 295f, 298
schematic model for the Sun 3013
solution of solar neutrino problem
38486
supernova core 333f
supernova mantle 43242
spin-precession picture 286f, 314f
supernova
cooling phase 43436
explosion mechanism 436f
overview 430f
prompt burst 43234, 496f, 498
r-process nucleosynthesis 43742
Sun
MSW solutions 38486, 434f
schematic model of MSW eect
3013
vacuum solution 38184
survival probability
general expression 28486
with energy distribution 288
with source distribution 287f
temporal vs. spatial 283f
three-avor 284f
two-avor 28488
neutrino-photon scattering 245f
neutrino radiative decay
neutrino dipole moment, neutrino
two-photon coupling
Subject Index
neutrino radiative decay (contd)
bounds from
beam stop 456f
cosmic background neutrinos
48991
diuse ux from all stars 487f
reactors 45355
solar positrons 457, 461
solar x-rays 45961
SN 1987A -rays 467f, 474f, 47779
photon spectrum
stationary source 45153
pulsed source 46365, 46973
standard-model predictions 26467
summary of bounds 278f
neutrino rocket engine 44345
neutrino sphere 397, 407, 409
neutrino spin precession
supernova magnetic eld 521f
theoretical description
equations of motion 304f
in electric elds 305
in medium 306
oscillation length 305
spin-avor 3069
time variation of solar neutrino ux
374, 387f
neutrino trapping 396
neutrino two-photon coupling 245, 265f,
271f
neutrinos: cosmic background
screening of leptonic force 114f, 500
mass density 25860, 48991
radiative decay 48991
neutrinos: number of families 251, 260,
513
neutrinos: right-handed
component of Dirac 253f
e+ e decay limits 479, 482
SN 1987A limit
charge radius 523
dipole moment 521f
Dirac mass 51619
mixing with sequential avors
33840
right-handed currents 51920
spin-ip production 16064
neutrinos: solar solar neutrino flux
neutrinos: strongly interacting 558f, 563
657
neutrinos: supernova SN 1987A,
supernova core
neutron stars
cooling 5460
crust 5860, 106f
formation 38
kick velocities 44345
internal conditions 601
magnetosphere 18588
pseudoscalar eld around 187
x-ray observations 5658
Newtons constant
value 6
time variation 54654
NOMAD experiment 289f
nonabelian Boltzmann collision
equation 323f
nuclear matter
novel phases 15559
properties 151f, 6005
nucleon mass: eective 151f, 6025
nucleosynthesis big-bang nucleosynthesis, supernova (type II)
number fraction: denition 243
O
o-diagonal refractive index 322
Oklo natural reactor 546
one-pion exchange potential 119, 122,
124f, 141
opacity
-enhanced 78f
conductive 11, 22, 78
denition 11
neutrino 13235, 14951, 329f, 334f,
408, 51416
particle 11, 18f, 12735
radiative 18f
reduced 19
Rosseland average 11, 18f
solar 21, 345, 35355
tables 12
OPAL 12
open clusters 60, 76
OPE potential 119, 122, 124f, 141
optical activity birefringence
optical theorem 198
oscillation length 285f, 289, 29497, 305
658
oscillations (particles)
neutrino oscillations
axion-photon 17982, 185f, 18890
graviton-photon 191
inhomogeneous media 181, 442f
pion-photon 183
oscillations (stars) helioseismology,
stars: variable
P
pair annihilation 99f
paraphoton 82, 92f, 99
Pauli matrices 286
Peccei-Quinn
mechanism 52628, 532f
scale 166, 526f, 529, 535
peculiar velocity (neutron stars) 44345
permittivity dielectric permittivity
phase velocity 199, 244
photon
baryonic 114
charge limits 555f
decay plasma process
leptonic 11416
mass
plasma frequency
eective 208
vacuum 206, 555f
transverse 209, 214
refraction photon dispersion
splitting 184
weakly interacting 18
photon dispersion
approximate expressions 213f
external elds (vacuum birefringence)
1835, 187f, 190f, 554
general theory 20319
mixing with axions 180f
transverse mass 209, 214
photoproduction
Compton process
X-particles in HB stars 82, 429
pinching of neutrino spectra 410f
pion
one-pion exchange potential
Compton scattering 157
condensate 59, 15557, 243
coupling to nucleons 119, 151f, 531
Subject Index
pion (contd)
decay 162, 167f, 256
decay constant 166f, 527
mass 119, 124f, 527
mixing with axion 527f, 534
Nambu-Goldstone boson 165f
Planck mass: denition 6
Plancks constant 580
planetary nebulae 36, 43, 51
plasma
conditions in stars 59199
coupling parameter 224, 59799
crystallization 46, 598
uctuations 17274
frequency 198, 211f, 596, 599
one-component 22325
strongly coupled 223f
two-component 223
plasma process
neutrino dipole moments
astrophysical impact
HB star lifetime 8082
neutron-star crust 59f
red-giant core mass 8387
supernova core 523
general theory 22736
gluonic 159
plasmino 195
plasmon
coalescence 17072
e+ e decay 195, 211
longitudinal 17072, 195, 202f, 213
decay plasma process
p-mode helioseismology
Poincare sphere representation 286
polarimetry of radio sources 190f
polarization tensor 20512, 230f, 238
polarization vector
neutrino avor
denition 286f
degree of coherence 313
individual modes 317
shrinking by collisions 314f
transverse 314, 325
plasma excitations 206f
positron ux: interplanetary 461, 487f
pp-chain 15, 341, 346f
pressure sources in stars 710
Subject Index
Primako process 116, 159, 16875,
177, 357
prompt
explosion mechanism 399, 401f
neutrino burst 397, 399f, 43234,
496f, 498
proton
decay xvii
fraction in nuclear matter 601f
spin 537, 584
protoneutron star supernova core
pseudomomentum 200
pseudo Nambu-Goldstone bosons 166
pseudoscalar bosons
arions, axions, majorons
astrophysical impact and bounds
HB star lifetime 8082, 98f, 176,
191f
neutron-star cooling 59
red-giant core mass 85f, 107-9
SN 1987A 5014, 50812
Sun 16f, 20f, 176, 191f
white-dwarf cooling 5052
ZZ Ceti period decrease 54
bounds on couplings to
photons 17, 81f, 176, 191f
electrons 20f, 51f, 82, 85, 99f,
10711
nucleons 512
conversion to photons 17982, 185f,
18890
derivative coupling 119, 531f, 563
emission processes
bremsstrahlung (electronic) 1025,
107f
bremsstrahlung (nucleonic) 11926,
14345, 148f, 155f
bremsstrahlung (quarks) 157f
Compton 20, 94, 98, 108f
energy spectrum 123, 175
pair annihilation 99f
Primako (gluonic) 159
Primako (photonic) 16875, 177
interaction Hamiltonian 119
long-range force 112
opacity contribution 18f, 21, 131f,
5068
refractive index 250
solar direct search 100, 181, 191f
659
PSR 0655+64 547
PSR 1913+16 (Hulse-Taylor binary
pulsar) xiv, 113, 116, 547
PSR 1937+21 188, 190, 554
pulsar
neutron star
pulse dispersion 554f
pulsation
constant 40
periods 40f
period decrease 53f
PVLAS experiment 185, 192
Q
QCD: CP problem 52426
Q-nuclei (quarked nuclei) 557
quark
free 556f
matter xv, 15759, 557
masses 252, 525
mass ratios 535
mixing 261f
quantum gravity 532f
R
radiative energy transfer energy
transfer, opacity
random media: particle oscillations 181,
442f
rank-ordering statistics 374f
reactor: nuclear
natural 546
neutrino radiative decay limits 454f
neutrino spectrum 450, 453f, 455
red-giant
branch RGB
core mass at helium ignition
increase by particle emission 8387
observational 7579
theoretical 67f
uncertainties 78f
core rotation 78
properties and evolution 2832, 592,
599
reduced opacity 19, 131
660
refractive index
photon dispersion, neutrino
dispersion
denition 194
forward scattering 19698
neutrino 242, 244
relativistic limiting velocity 497f
renormalization: wave function 200,
21719, 228, 233
repulsion of energy levels 296
resonant oscillations
neutrino oscillations
axion-photon 181, 186
RGB (red-giant branch)
brightest star 63f, 6872
bump 61, 64f
location in color-magnitude diagram
27, 41
phase transition 37
right-handed
currents 268, 51920
neutrinos neutrinos: right-handed
Ring Nebula in Lyra 36
R-method 65, 70, 72, 7478
ROSAT 5658
Rosseland average 11, 18f
r-process nucleosynthesis 4056, 43742
RR Lyrae stars
absolute brightness 62, 65, 69f, 73f
ages of globular clusters 62
color 62
location in color-magnitude diagram
40f
mass-to-light ratio 77
Rubakov-Callan eect xv
Rydberg atom 178
S
S17 factor 35557, 377f, 386
SAGE 343f, 35759, 362f, 371
Sanduleak 69 202 39, 414
scalar eld
scalar particles, Higgs field
around a neutron star 113
scalar particles
bounds on coupling
electrons 82, 99, 107
nucleons 82, 113
Subject Index
scalar particles (contd)
emission from binary pulsar 113
emission processes
bremsstrahlung 103, 124f
Compton 93, 98
long-range force 112f, 500
photon coupling 166
Sciamas neutrino 491f
screening eects
background neutrinos 114f
Debye-H
uckel scale 220f, 596f, 599
general theory 21926
in processes
bremsstrahlung 1026
Primako 169
spin-ip scattering 277f
ion contribution 220f, 596f, 599
modication of Coulomb propagator
221f, 226
Thomas-Fermi scale 221
secret neutrino interactions 558f, 563
see-saw mechanism 254
self-energy 201, 209
Shapiro time delay 498, 554
shining light through walls 182
shock wave 397, 399405
Small Magellanic Cloud 446, 556
SMM satellite 465f
SN supernova
SN 1987A
neutrino pulse
analysis 42326
anomalies 42730
measurements 41923
optical lightcurve 414f
progenitor 39, 414
SN 1987A bounds
antimatter supernova 497
arion-photon conversion 189
axions
direct detection 495
neutrino signal duration 5014,
50812
fundamental length scale 500
Lorentz addition of velocities 500f
neutrino
charge 499, 522f
decay 494, 496, 519
mass 426f, 499
Subject Index
SN 1987A bounds (contd)
neutrino (contd)
neutrino-neutrino cross section 494f
number of families 513
secret interactions 494f
neutrino oscillations
cooling phase 43436
prompt burst 43234
pseudoscalar boson couplings 5014,
50812
radiative particle decays 467f, 474f,
47779, 483
relativistic limiting velocity 497f
right-handed currents 51920
right-handed neutrinos
charge radius 523
dipole moment 521f
Dirac mass 51619
mixing with sequential neutrinos
33840
secret neutrino interactions 563
supersymmetric particles 558
weak equivalence principle 498
SNBO 448
SNO 343, 371, 392f
SNu (supernova unit) 488
SNU (solar neutrino unit) 358
solar axions 100, 181f, 191f
solar maximum mission satellite 465f
solar neutrino ux
antineutrino component
from majoron decay 389
from spin-avor oscillations 388
limits from Kamiokande 369
counting rate prediction for detection
Cherenkov 370
chlorine 359, 361
gallium 359, 364
future experiments 3904
measurements
Cherenkov 36870
chlorine 36062
gallium 36264
modied by
electrically charged neutrinos 565
gravitational settling 14, 352, 354
opacities 35355
neutrino decay 389
neutrino-neutrino scattering 495
661
solar neutrino ux (contd)
modied by (contd)
resonant oscillations 3013, 38486,
434f
temperature 354
WIMP energy transfer xvi
strange quark matter 557
Q-nuclear burning 557
time-varying GN 54951
vacuum oscillations 38184
radiative decay limits 45862
source reactions
beryllium 343, 34750, 355, 378f
boron 343, 34749, 35557, 36870,
377f, 383
CNO 343, 34749
electron capture vs. decay 350
hep 34749
pep 347, 350
pp 343, 34749
time variation
day-night 372, 385f, 390f
semiannual 372f, 382f
solar cycle 37376, 387f
solar neutrino problem
introduction and historical overview
34145, 380f
ux decits
beryllium 378f
beryllium/boron branching ratio
379f
boron 377f
ux variation at Homestake 37377
MSW solution 38486
vacuum solution 38184
VVO solution (magnetic oscillations)
387f
Sommerfeld parameter 355
space-like excitations 194, 198f, 207,
216, 215f, 238f
spectral density 173f
speed of light 497f
spin-avor oscillations neutrino spin
precession
spin ip 16064, 277f, 3049, 317,
51623
spin-uctuation rate 118, 121-23, 127,
133, 144f
spin relaxation 313
662
spin-spin interaction potential 141, 151
starburst galaxies 446
stars
ages 27, 35, 43, 62f, 55254
formation 2427
initial mass function 24
intermediate-mass 37
mass loss 25, 34, 36
mass range 24
massive 3739
populations in color-magnitude
diagram 41
variable 12, 3941, 5254
statistical parallaxes 73
Stefan-Boltzmann law 409
stellar collapse supernova: type II
stellar evolution
bibliography 24
descriptive overview 2341
evolutionary track 31
main phases 30
stellar oscillations helioseismology,
variable stars, ZZ Ceti stars
stellar structure
convective 12f
equations 514
examples for models 29f
generic cases 710
homologous models 1416, 549, 552
long-range force: new 113
sterile neutrinos
neutrinos: right-handed
Stodolskys formula 31315, 324f
Stokes parameters 286
strange quark matter xv, 15759, 557
structure function
Coulomb plasma 106, 22225
dynamical
classical 146f
detailed balance 129, 137, 146
formal denition 13643, 323
long-wavelength limit 121, 128f,
14143
Lorentzian model 135, 14447
nuclear medium 128f, 135, 14651,
16163
spin-density 138
static 137, 22225
subgiant 27, 28
Subject Index
sum rules 13941
Sun
solar neutrino flux
activity cycle 37376, 387f
axion spectrum 175
bounds on
axion ux 100, 181, 191f
time-varying GN 54951
deection of neutrinos 247f
energy-loss argument 16f, 20f, 109f,
175f
global properties
central temperature 7, 353
convective layer 550
distance 341, 372
Kelvin-Helmholtz time scale 8
helium abundance 16f, 351, 549f
luminosity 8, 341
magnetic eld 186, 37376, 387f,
554, 565
mass 7
plasma properties 599
radius 7, 351
opacity 21, 345, 35355
positron ux limits 457, 461
spots 186, 37376
standard model 29, 35153
x- and -ray ux 45862
superuidity in neutron stars 58f
Superkamiokande (Cherenkov detector)
343, 371, 3902
supernova
SN 1987A
burst observatory 448
core collapse supernova: type II
energetics: particle bounds 483
future 44548
Keplers 39
galactic
positron ux 484f
rate 39, 446, 487f
remnants 38f, 5658
SN 1054 38
Tychos 39
type I 36, 546
unit (SNu) 488
Subject Index
supernova core
binding energy 407
characteristics of the medium 398,
505f, 512, 6025
matter accretion 397, 400, 425f, 503
neutrino cooling
SN 1987A
analytic emission models
expected detector signal 418
schematic picture 397, 400, 407f
spectral characteristics 40811
time evolution 41114
neutrino avor conversion 33238
particle cooling
axion emission 5014, 50812
general argument 5018
numerical studies 50811, 51316
structure (numerical model) 505f, 512
supernova: type II
supernova core
description 3739
explosion mechanism 4015, 436f, 522
neutrino oscillations
cooling phase 43436
explosion mechanism 436f
overview 430f
prompt burst 43234
r-process nucleosynthesis 43742
nucleosynthesis 4056, 43742
stellar collapse 39599
supersymmetric particles
dark matter xvxvi, 22, 558
emission from stars 81, 557f
supersymmetry: avor-changing neutral
current 303
SXT satellite 186
T
tau neutrino
charge 567
dipole moment 278, 455, 457, 476f
e+ e decay: bounds
galactic supernovae 48486
reactors 455, 457
SN 1987A -rays 48082
solar positrons 457, 461
mass 256, 25860, 485f
theoretical decay rate 264
663
thermal broadening of beryllium line
350f
thermal equilibrium in stars 8
Thomas-Fermi scale 221
transition moment neutrino dipole
moments
transverse current 204, 219f
transverse gauge 204
transverse part of the avor
polarization vector 314, 325
transverse photon mass 209, 214
triangle condition 153f
triangle loop 167f, 530, 536
triangle: MSW 3013, 386, 434
triple- reaction 33, 80
tritium decay 255
twisting magnetic eld 308f
two-photon coupling
neutrinos 245, 265f, 271f
various bosons 16568
U, V
units
conversion factors 58082
natural 6
rationalized 184, 580
URCA process 1, 58f, 15254
vacuum birefringence 1835, 187f, 190f,
554
vector bosons
Compton process 9193, 98
energy-loss bounds 82, 99, 11012
long-range force 11216, 500
virial theorem
introduction 6f
negative specic heat 7f
VVO eect 387
W
weak damping limit (of neutrino
oscillations) 326f, 330f
weak decay spectrum 348
weak mixing angle 583
Weinberg angle 583
664
white dwarfs
bounds on
neutrino dipole moments 4850
pseudoscalar bosons 5052
time-varying GN 551
characteristics 9, 4245, 599
cooling theory 4547
formation 35f
inferred galactic age 43, 51
location in color-magnitude diagram
40f
luminosity function 4345
magnetic 185f
mass-radius relationship 9, 42
neutrino cooling 4750
variable 41, 5254
vs. red giant core 88
Subject Index
Whole-Earth Telescope 53
WIMP xv, 22
X, Y, Z
x-rays
neutron stars 5658
particle bounds
SN 1987A arion-photon conversion
189
SN 1987A axion decays 483
SN 1987A radiative neutrino
decays 467f, 474f, 47779
solar x-rays 45862
satellites 5658, 186, 465f
Yohkoh satellite 186
ZZ Ceti stars 41, 5254