You are on page 1of 7

Neutrino oscillations: A relativistic example of a two-level system

Elisabetta Sassaroli
Laboratory for Nuclear Science and Department of Physics, Massachusetts Institute of Technology,
Cambridge, Massachusetts 02139 and Department of Physics, Northeastern University, Boston,
Massachusetts 02115

Received 30 April 1998; accepted 3 March 1999


Neutrino flavor oscillations are discussed in terms of an explicit model. This model consists of two
coupled Dirac equations with three parameters: the electron neutrino mass, the muon neutrino mass,
and a coefficient which describes the possibility that the neutrino can flip flavor. The system is
diagonalized to obtain the exact eigenvalues and eigenfunctions. The system is then quantized and
the neutrino flavor wave functions are derived directly from the quantized fields. It is shown that
neutrino flavor oscillation probabilities are recovered in a quantum field theory treatment only in the
ultrarelativistic limit. 1999 American Association of Physics Teachers.

I. INTRODUCTION
Neutrino flavor oscillations are a relativistic example of a
two-level system. In his Lectures on Physics, Feynman1 describes many examples in which the approximation of a twostate system can be assumed. Some of his examples include
the ammonia molecule, the hydrogen molecule, a spin 1/2
particle in a magnetic field, and oscillations of strangeness in
the neutral K meson system.
For example, the ammonia molecule (NH3) has the form
of a pyramid with the nitrogen atom located above the plane
of the three hydrogen atoms. Like any other, this molecule
has an infinite number of states; however in the two-level
system approximation, it is assumed that all the states remain
fixed except for two: the nitrogen may be on one side of the
plane of the hydrogen atom or on the other. The system can
be described by the state vector ,
C 1 1 C 2 2 ,

where in the state 1 the nitrogen is up and in the state 2


the nitrogen is down. The coefficient C 1 1 is the
amplitude to be in state 1 and C 2 2 the amplitude to
be in state 2. The coefficients C 1,2 are obtained by diagonalizing the two coupled differential equations
i

dC 1
H 11C 1 H 12C 2 ,
dt

dC 2
H 21C 1 H 22C 2 ,
i
dt

with H i j being the Hamiltonian matrix, which depends on


the particular system which is studied. Feynman solved and
discussed in great detail the set of equations given by Eq. 2
for the ammonia molecule case and also applied these equations to the ammonia maser.
In this paper we will to consider a relativistic generalization of the above set of two coupled differential equations to
the neutrino flavor oscillation case.
Neutrinos are relativistic noncharged particles of spin 1/2,
which are produced in weak interaction processes.2,3 In the
electroweak theory of Glashow, Salam, and Weinberg
GSW, neutrinos are massless and they can exist in three
different flavors: the electron, muon, and tau flavors. However as far as we know, there is no deep theoretical reason
why neutrino masses, i.e., their rest energies, should be ex869

Am. J. Phys. 67 10, October 1999

actly zero. Moreover neutrinos have masses in most extensions of the GSW theory. It is therefore extremely important
to investigate experimentally and theoretically neutrino
masses.
The experimental investigation of the allowed direct and
inverse decay see, for example, Ref. 4 for more details
pne e ,
e pne ,

e npe ,
and forbidden processes

e pne ,

e npe ,

can be conveniently described by introducing an electron


flavor quantum number. The electron e and its neutrino e
are assigned the value L e 1, the positron e and the antineutrino e are assumed to have L e 1, while all the
remaining particles for example, the proton p and the neutron n in the above reactions are assigned the value L e 0.
The electron flavor number is conserved in the allowed processes the total electron flavor number on the left-hand side
of the reaction is equal to the total electron flavor number on
the right-hand side, while the forbidden ones would violate
this conservation law. Moreover the studies of processes
such as, for example,

e e ,
pn ,
and the fact that processes of the type

e ,

pne ,

are not observed experimentally lead to the introduction of


the muon flavor quantum number L 1 for muon and
its neutrino , L 1 is for to and , and L 0
for all the remaining particles. The tau leptons
( , , , ) are described with the help of the tau flavor
number L .
All the examples considered above satisfy the conservation of both lepton numbers L e and L separately the total
lepton number on the left-hand side of the reaction is equal
to the total lepton number on the right-hand side. Because
1999 American Association of Physics Teachers

869

the origin of these quantum numbers is not clear, there is the


possibility that they are only approximately conserved. If so,
then a neutrino produced in a given flavor can transform into
another flavor while propagating, as will be discussed in Sec.
II. If this is the case, the phenomenon of neutrino flavor
oscillations can arise in nature. Recent experiments strongly
suggest the evidence of neutrino oscillations.5 More specifically, neutrino flavor oscillations can be a possible explanation of the atmospheric neutrino anomaly measured by three
different experiments68 of the solar neutrino deficit observed by four different experiments,913 and of the evidence
of neutrino masses obtained by the liquid scintillator nuclear
detector LSND experiment.14 Neutrinos are also considered
to make a small contribution to the dark matter of the universe hot dark matter.
We point out here that the properties of neutrinos electric
charge zero, mass very small, maybe zero make their detection very difficult. A typical interaction cross section between an electron neutrino and a nucleus at 1 MeV is of the
order of 1043 1044 cm2.
In this paper we are going to study neutrino oscillations as
an example of a two-level system. The two coupled differential equations, which describe two-level systems, are valid
in the rest frame of the system investigated. However, since
neutrinos are relativistic particles of spin 1/2, it is important
to write a relativistically invariant generalization of Eq. 2.
The simplest generalization of Eq. 2 which is relativistically invariant and properly takes into account the neutrino
spin is a system of two coupled Dirac equations with three
parameters: the electron neutrino mass, the muon neutrino
mass, and a coefficient which describes the possibility that
the neutrino can flip its flavor, as will be discussed in Sec.
III. For simplicity we are only investigating a two-flavor
model. The system is diagonalized and the energy eigenvalues and eigenfunctions are obtained. These solutions represent an interesting new piece of relativistic quantum mechanics. Contrary to the nonrelativistic case where the
diagonalization of the two coupled equations above gives
directly the probability amplitudes for flipping from one state
to the other, in a relativistic system, in order to deal properly
with the states of negative energies we have to abandon the
one-particle picture and adopt a many-particle formulation,
e.g., quantum field theory.
The solutions of the two-coupled Dirac equations are
quantized according to the usual JordanWigner anticommutation relations and the neutrino energy wave functions are
obtained as matrix elements of the quantized Dirac fields, as
described in Sec. IV. These wave functions describe neutrinos of given energy and in a state of mixed flavor at any
spacetime point. It is also shown that their sum describes
neutrinos in a state of mixed flavor. Therefore it is not possible with Dirac fields to impose the boundary condition to
have a given flavor for all the space points at a given time,
lets say at production. Hence, Dirac fields cannot properly
describe neutrino oscillations, where it is assumed that the
neutrino is produced in a state of given flavor and then subsequently oscillates in flavor. Neutrino oscillation probabilities can be recovered only if the following two approximations are made on the neutrino flavor wave function: i the
left-handed chiral component of the flavor wave function is
considered as an observable wave function; ii the ultrarelativistic approximation is assumed for the spinor component
of the wave function.
Approximation i is due to the fact that neutrinos are
870

Am. J. Phys., Vol. 67, No. 10, October 1999

produced only through the weak interaction, which does not


conserve parity. Approximation ii appears to be related to
the impossibility of simultaneously maintaining Lorentz invariance and obtaining standard neutrino oscillation probabilities in a way which is consistent with relativistic field
theory. This problem is obviously a deep one and is associated with the possibility that neutrinos violate Lorentz invariance as well as the equivalence principle as discussed in
Refs. 1517.

II. NEUTRINO OSCILLATIONS: STANDARD


TREATMENT
We will first review the standard quantum mechanical
treatment of neutrino flavor oscillations. Suppose for example that a muon neutrino is produced in the reaction

,
and while it is propagating has a probability of flipping its
flavor and becoming an electron neutrino, in the same way as
it is possible for the nitrogen atom in the ammonia molecule
to push its way through the three hydrogen atoms and flip to
the other side, due to a quantum tunneling effect.
The system can be described by a state vector as a
linear combination of the flavor eigenstates e and ,
C e e C ,

Ce
,
C

with C e e , C , and
C e 2 C 2 1.

C e and C then become the amplitudes for detecting an electron neutrino and a muon neutrino, respectively. In analogy,
in the neutral K meson system the oscillations occur between
0 (S
states of different strangeness K 0 (S1) and K
1).
To derive the time evolution of the coefficients C e (t) and
C (t), the state vector is written as a superposition of the
energy mass eigenstates I and II ,
C I 1 C II 2 ,

CI
,
C II

with C 1 I , C 2 II , and
C I 2 C II 2 1,

where C I and C II are the amplitudes for finding the neutrino


in the energy states E 1 and E 2 , respectively. These coefficients evolve in time as
C I t C I 0 e iE 1 t ,

C II t C II 0 e iE 2 t .

Introducing the rotation matrix between flavor and mass


eigenstates

cos
I

II
sin

sin
cos

e
,

10

it is easy to see that the following relation between the energy and flavor amplitudes holds:
Elisabetta Sassaroli

870

cos
C I t

C II t
sin

sin
cos

C e t
.
C t

11

Hence, the time evolution of the coefficients C e and C is


given by

cos
C e t

C t
sin

sin
cos

C I 0 e iE 1 t
.
C II 0 e iE 2 t

12

In Eq. 12 the boundary condition must be imposed that


we have only a given flavor at production. Suppose for example that at t0 a muon neutrino is produced, i.e.,
C 0 1,

C e 0 0.
C 2 0 cos .

14

C e t sin cos e iE 2 t e iE 1 t ,

15

iE 1 t

16

C t sin e

cos e
2

iE 2 t

Space and therefore momentum is introduced by assuming


in Eqs. 15 and 16
E 21 m 21 p 2 ,

E 22 m 22 p 2 ,

17

Lt.

The probability of finding a given flavor is obtained by


squaring Eqs. 18 and 19,
C e 2 sin2 2 sin2 E 2 E 1 t/2

sin 2 sin
2

m 21 m 22 L

4E

18

1sin 2 sin

m 21 m 22 L

4E

19

with Ep. In order to compare with experimental data, the


probabilities given by Eqs. 18 and 19 have to be averaged
over the energy distribution of particles involved. For a more
complete analysis of the phenomenology of neutrino flavor
oscillations see, for example, Ref. 3.
The assumption that the muon neutrino is created with a
definite momentum p is only an approximation, as has been
pointed out previously.1821 It is in contradiction with fourmomentum conservation, for example for the reaction
. Each of the possible energy eigenstates has a somewhat different momentum pi . In the rest frame of the pion,
energy conservation dictates that (i1,2)
M M 2 p2i m 2i p2i .

20

III. DIAGONALIZATION OF THE TWO COUPLED


DIRAC EQUATIONS
In this section we will derive a relativistic generalization
of the two coupled equations, given by Eq. 2, in the context
of the Dirac theory. The system of equations is diagonalized
and the energy eigenvalues and eigenfunctions are derived.
Suppose, for example, that a muon neutrino is produced
with some small mass rest energy. The propagation of the
871

Am. J. Phys., Vol. 67, No. 10, October 1999

21

e x,t
p m e e x,t .
t

22

Now suppose that, due to a quantum tunneling effect, the


muon neutrino has some probability of flipping its flavor and
turning into an electron neutrino. If is the parameter that
describes the flavor flipping possibility, then neutrino flavor
oscillations can be described by the two coupled Dirac equations
i

e x,t
p m e e x,t x,t ,
t

23

x,t
p m x,t e x,t .
t

24

We notice here that has the dimension of a mass. The


details of the diagonalization of the system of equations are
given in the Appendix; here we simply discuss the solutions.
For a given momentum p p the energy eigenvalues are
2
E 1,2 p 2 m 1,2
,

25

where m 1,2 are the renormalized masses

C 2 1sin2 2 sin2 E 2 E 1 t/2


2

x,t
p m x,t ,
t

with p i and and are the Dirac matrices.


If the muon neutrino does not undergo any flavor change,
then it will move as a free particle according to Eq. 21. In
the same way, if a neutrino is produced as an electron neutrino and has a small mass, but it does not oscillate in flavor,
its propagation in vacuum will be described by
i

The time evolution of the flavor amplitudes is obtained by


substituting Eq. 14 into Eq. 12,
2

13

From Eq. 11 at t0 we obtain


C 1 0 sin ,

free neutrino can be described by the Dirac equation for a


review about the Dirac equation and its solutions see, for
example, Ref. 22

m 1,2 21 m e m R ,

R m m e 2 4 2 .

26

A system of two coupled Dirac equations possesses for a


given value of the energy two eigenfunctions; one with spin
up and one with spin down. If 1 (x,t) is the eigenfunction
corresponding to the positive energy solution E 1 given by
Eq. 25, it is possible to show that it can be written in terms
of a two-dimensional vector

Z 1

1 2

m m e R
2

27

1 2

times the solution of the Dirac equation of renormalized


mass m 1 , momentum p, and energy E 1 ,

1 x,t Z 1

V 2E 1

u 1 s,p e ipx e iE 1 t ,

28

where s1,2 is the spin index, u 1 (s,p) is the Dirac spinor of


mass m 1 .
The wave function 1 (x,t) has eight dimensions because
the two-dimensional vector Z 1 multiplies a four-dimensional
wave function. This last wave function describes a neutrino
of given energy E 1 and spin 1/2. The vector Z 1 tells us that
Elisabetta Sassaroli

871

in any location inside the volume V there is a probability


equal to (1/1 2 ) of finding the neutrino in the electron
flavor and probability equal to ( 2 /1 2 ) of finding it in
the muon flavor. Therefore 1 describes a neutrino of a
given energy, but of mixed flavor at any spacetime point.
In the same way, the wave function corresponding to the
other positive energy solution E 2 is

2 x,t Z 2

V 2E 2

u 2 s,p e ipx e iE 2 t ,

29

where Z 2 is the vector

Z 2

1 2

30

and u 2 (s,p) is the Dirac spinor of renormalized mass m 2 .


The wave function 2 describes a neutrino of given energy E 2 and mixed flavor at any spacetime point.
The solutions of negative energies are interpreted, as in
the Dirac theory, as antiparticles of positive energy and are
given by
1

x,t

V 2E 1,2

v 1,2 s,p e ipx e iE 1 t ,

The quantization procedure for the two coupled Dirac neutrino fields e and , defined by Eqs. 23 and 24, proceeds in the same way as for the case of Dirac theory. For a
discussion of the quantization of the Dirac theory, see, for
example, Ref. 22.
We expand the neutrino field in terms of the energy
eigenfunctions and found in Sec. III,

e x,t
x,t

32

where the operators b i and d i (i1,2) satisfy the usual


JordanWigner anticommutation relations

b i s,p ,b j s ,p i j pp ss ,
d i s,p ,d j s ,p i j pp ss .

33

For a given value of the momentum p and spin s, there are


four possible one-particle states, one for each energy value,
given by Eq. 25

d 1 s,p 0 1 ps ,
872

36

A 2 B 2 1.

37

The matrix element

0 e x,t e x,t

V 1

u 2 s,p

2E 2

u 1 s,p

2E 1

e iE 1 t

e iE 2 t e ipx ,

38

gives the probability amplitude of finding a neutrino of momentum p and spin s at the spacetime point (x,t) with the
electron flavor. In the same way, the matrix element

0 x,t x,t

V 1
B

u 2 s,p

2E 2

u 1 s,p

2E 1

e iE 1 t

e iE 2 t e ipx ,

39

is the probability amplitude for the muon flavor.


The coefficients A and B are determined through the initial
boundary conditions. Suppose that at t0,

p s i b i s,p i x,t d i i x,t ,

b 1 s,p 0 1 ps ,

35

where A and B satisfy the normalization condition

31

IV. FIELD QUANTIZATION, ANTICOMMUTATION


RELATIONS, AND FLAVOR WAVE FUNCTIONS

It is easy to see that the above wave function is the energy


eigenfunction of energy E 1 defined in Eq. 30. Similar considerations can be applied to the other states 2 ps , 1 ps ,
and 2 ps . These one-particle states represent states of
mixed flavor at any given spacetime point.
When describing neutrino oscillations, we have to consider a superposition of energy states. A general state of
positive charge, momentum p, and spin s is given by

where v 1,2(s,p) are Dirac spinors of mass m 1 and m 2 , respectively.

x,t

e x,t
0 e x,t 1 ps

.
x,t
0 x,t 1 ps

Ab 1 s,p Bb 2 s,p 0 ,

1 2

1,2 x,t Z 1,2

The wave function associated with the one-particle state


1 ps is, for example, obtained as a matrix element of neutrino fields between the vacuum state and the one-particle
state,

b 2 s,p 0 2 ps ,
d 2 s,p 0 2 ps .

Am. J. Phys., Vol. 67, No. 10, October 1999

34

x,t0 0,

40

i.e., we have only the electron flavor present. The other


boundary condition is obtained from the normalization condition

d 3 x e x,t0 2 1.

41

However, the above boundary conditions cannot be applied in a consistent way to the flavor wave functions given
by Eqs. 38 and 39 and at the same time satisfy the conservation of probability condition given by Eq. 37. The
following two approximations have to be made on the neutrino wave functions in order for us to be able to impose the
boundary conditions given by Eqs. 40 and 41.
Elisabetta Sassaroli

872

a The left-handed chiral components of the flavor wave


functions are considered as observable wave functions.
Mathematically this is equivalent to considering as observable wave functions23

eL x,t 1 5 e x,t ,

L x,t 1 5 x,t ,
42

where e (x,t) and (x,t) are given by Eqs. 38 and 39


and 5 is defined by

43

b The ultrarelativistic approximation, i.e., E p, is assumed in the spinor left-handed chiral components u L (s,p)
of the flavor wave functions.
By applying the approximations a and b one obtains
the flavor neutrino wave functions

eL x,t

L x,t

1
1 iE t
2
,
e 1 2 e iE 2
2
2
V 1 &
44

e ipx

1 iE t iE t 2
.
e 1 e 2
2
2
V 1 &
45

e ipx

The probability densities of finding the electron and muon


neutrino flavor are then given, respectively, by

2
1
e t 1
V
1 2

2
1
V 1 2

sin2

E 2 E 1 t
sin2
,
2

46

E 2 E 1 t
.
2

47

The coefficient 2/(1 2 ) 2 is equivalent to sin2(2) in


Eqs. 18 and 19. Therefore Eqs. 46 and 47 are equivalent to the standard neutrino oscillation probabilities.
It is important to discuss the meaning of both approximations. Approximation a takes into account the fact that neutrinos are produced through the weak interaction, which does
not conserve parity. Therefore neutrinos are created with
negative helicity and antineutrinos with positive helicity, i.e.,
neutrinos are emitted with their spin polarized opposite to
their direction of motion and antineutrinos have their spin
polarized in the same direction. This statement is exactly true
on the hypothesis that neutrinos are massless and it is certainly a very good approximation for very small neutrino
masses.
Approximation b is related to the conservation of probability. The coefficients A and B, obtained by imposing the
boundary conditions given by Eqs. 40 and 41 in the flavor
wave functions eL (z,t), L (z,t), do not satisfy the normalization condition given by Eq. 37. This condition is
satisfied only if we assume that the terms of type p/(E
m) in spinors u L are of order one, i.e., in the ultrarelativistic limit. Therefore only in the ultrarelativistic limit can we
impose the condition of a given flavor at production without
violating the condition that the sum of the probabilities must
be one.
Other authors,24,25 by using different approaches, have
also found that the standard neutrino oscillations can be recovered only in the ultrarelativistic limit.
873

Am. J. Phys., Vol. 67, No. 10, October 1999

We believe that there are deep theoretical reasons for the


impossibility of discussing in a consistent way flavor oscillations in field theory. In the literature, some hypotheses
have been proposed in relation to this issue, especially in
relation to the problem of CPT and Lorentz violations. For
example, Coleman and Glashow have examined the assumption that Lorentz noninvariance leads to neutrino oscillations
which are phenomenologically equivalent to those obtained
by assuming that neutrinos violate the equivalence principle.
A variety of different approaches have been considered in
the literature to address the problem of CPT and Lorentz
invariance violations, as discussed in Ref. 26.
V. CONCLUSIONS
We have investigated an explicit model of neutrino flavor
oscillations in the framework of relativistic quantum mechanics and quantum field theory. This model, which is a
relativistically invariant generalization of a two-level system,
consists of two coupled Dirac equations. The system has
been diagonalized and second quantized in order to deal
properly with the states of negative energy. The neutrino
wave functions are obtained as matrix elements of the quantized neutrino fields. These wave functions, however, describe neutrinos which are in a state of mixed flavor at any
spacetime point and only in the so-called ultrarelativistic
limit do they describe the possibility of having a given flavor
at production.
ACKNOWLEDGMENTS
This work is supported in part by funds provided by the
U.S. Department of Energy D.O.E. under cooperative research agreement No. DF-FC02-94ER40818.
The author would like to thank Professor Alan H. Guth for
his constructive criticism which led to a revision of the paper. She would also like to acknowledge fruitful discussions
with Professor Kenneth Johnson and the MIT atomic and
molecular interferometry group.
APPENDIX
In order to determine the energy eigenvalues and eigenfunctions of the system of equations 23 and 24 we consider the ansatz

e ae i Px ,

A1

i Px

A2

be

where P is the four-momentum P(E,p), which is unknown and is to be determined so that the system of differential equations 23 and 24 is satisfied. The coefficients a
and b are Dirac spinors, which can be written as
a
b

1
,
2

A3

3
,
4

A4

where 1,2 and 3,4 are two component spinors. Substituting


Eqs. A1 and A2 into Eqs. 23 and 24, we obtain the
system of linear homogeneous equations
E 1 p 2 m e 1 3 ,

A5

E 2 p 1 m e 2 4 ,
Elisabetta Sassaroli

873

E 3 p 4 m 3 1 ,

u 2 s,p E 2 m 2

A6

E 4 p 3 m 4 2 ,

where are the Pauli matrices.


The system of Eqs. A5 and A6 admits nontrivial solutions only if

A7

E 1,2 p

A8

with m 1,2 given by


m 1,2 21 m e m R ,

A9

R m m e 2 4 2 .

A10

Therefore, while in the free Dirac equation there are two


energies one positive and one negative for every possible
value of the momentum p, for a system of two coupled Dirac
equations, there are four possible values of the energy, two
positive and two negative. This is due to the possibility of
flavor oscillations. Also, because there is some chance that
the neutrino can flip flavor, the rest energies of the electron
and muon neutrino system are not simply m e and m but are
given by Eq. A9.
Corresponding to the positive energy solution E 1
m 21 p 2 , we have the following two solutions:
1

V 2E 1

1 s,p e ipx e iE 1 t ,

A11

1 x,t

u 1 s,p
,
u 1 s,p

A12

and u 1 (s,p) is the Dirac spinor


u 1 s,p E 1 m 1

s
p
,
s
E 1 m 1

3 s,p

m m e R
.
2

with 2 (s,p) given by

2 s,p

A15

A16

u 2 s,p
,
u 2 s,p

and the Dirac spinor u 2 (s,p) is


874

Am. J. Phys., Vol. 67, No. 10, October 1999

u 2 s,p
.
u 2 s,p

A19

V 2E 1

1
1

3 s,p e ipx e iE 1 t ,

v 1 s,p
v 1 s,p

v 1 s,p
,
v 1 s,p

A20

A21

p
s
E 1 m 1
.
s

A22

For the energy eigenvalue E 2 we have the solution

2 x,t

V 2E 2

with 4 (s,p) given by

A13

2 s,p e ipx e iE 2 t ,

V 2E 2

and

A14

v 1 s,p E 1 m 1

4 s,p

For the other positive energy solution E 2 p 2 m 22 , we


have

2 x,t

and

v 2 s,p E 2 m 1

with

with 3 (s,p) given by

where s1,2 is the spin index and 1 (s,p) is given by

1 s,p

A18

Similarly, for the solutions of negative energies E 1 we


have the eigenfunction:

and

1 x,t

m m e R
.
2

2 s,p

2
m 1,2
,

A17

We notice here that because 1 we can write


2 (s,p) in terms of as

Solving Eq. A7, we obtain (p p )


2

is defined as

E 4 E 2 2p 2 2 2 m 2e m 2 p 4
4 p 2 2 2 m 2e m 2 m 2 m 2e 2 2 m e m 0.

s
p
.
s
E 2 m 2

4 s,p e ipx e iE 2 t ,

A23

A24

v 2 s,p
,
v 2 s,p

p
s
E 2 m 2
.
s

A25

R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman Lectures in


Physics AddisonWesley, Reading, MA, 1965, Vol. III.
2
For a review of neutrino physics, see for example: S. M. Bilenky and B.
Pontecorvo, Phys. Rep., Phys. Lett. 41C, 225 1978; F. Boehm and P.
Vogel, Physics of Massive Neutrinos Cambridge U.P., Cambridge, England, 1987; B. Kayser, F. Gibrat-Debu, and F. Perrier, The Physics of
Massive Neutrinos World Scientific, Singapore, 1989; J. N. Bahcall,
Neutrino Physics and Astrophysics Cambridge U.P., Cambridge, 1989;
R. N. Mohapatra and P. B. Pal, Massive Neutrinos in Physics and Astrophysics World Scientific, Singapore, 1991.
3
C. W. Kim and A. Pevsner, Neutrinos in Physics and Astrophysics Harwood, Chur, 1993.
4
D. H. Perkins, Introduction to High-Energy Physics AddisonWesley,
Reading, MA, 1982.
5
Super-Kamiokande Collaboration, Y. Fukuda et al., Evidence of oscillations of atmospheric neutrinos, Phys. Rev. Lett. 81, 15621567 1998;
hep-ex/9807003.
Elisabetta Sassaroli

874

K. S. Hirata et al., Experimental study of the atmospheric neutrino


flux, Phys. Lett. B 205, 416420 1988.
7
R. Becker-Szendy et al., The electron-neutrino and muon neutrino content of the atmospheric flux, Phys. Rev. D 46, 37202724 1992; D.
Casper et al., Measurement of atmospheric neutrino composition with
the IMB-3 detector, Phys. Rev. Lett. 66, 25612564 1991.
8
W. W. M. Allison et al., Measurement of the atmospheric neutrino flavour composition in Soudan 2, Phys. Lett. B 391, 491500 1997.
9
B. T. Clevend et al., Update on the measurement of the solar neutrino
flux with the Homesteke chlorine detector, Nucl. Phys. B Proc. Suppl.
38, 47 1995; R. Davis, A review of the Homestake solar neutrino
experiment, Prog. Part. Nucl. Phys. 32, 1332 1994.
10
K. S. Hirata et al., Constraints on neutrino oscillation parameters from
the Kamiokande-II solar neutrino data, Phys. Rev. Lett. 65, 13011304
1990; Real time, directional, measurement of B-8 solar neutrinos in the
Kamiokande-II detector, Phys. Rev. D 44, 22412260 1991.
11
Y. Fukuda et al., Solar neutrino data covering solar cycle 22, Phys.
Rev. Lett. 77, 16831686 1996.
12
J. N. Abdurashitov et al., Results from Sage, Phys. Lett. B 328, 234
248 1994.
13
P. Anselmann et al., Gallex results from the first 30 solar neutrino runs,
Phys. Lett. B 327, 377385 1994; 342, 440 1995.
14
C. Athanassopoulos et al., Evidence for neutrino oscillations from muon
decay at rest, Phys. Rev. C 54, 26852708 1996; nucl-ex/9605001.
15
S. Coleman and S. L. Glashow, Cosmic ray and neutrino tests of special
relativity, Phys. Lett. B 405, 249252 1997.
16
S. L. Glashow, A. Halprin, P. I. Krastev, C. N. Leung, and J. Pantaleone,

Comments on neutrino tests of special relativity, Phys. Rev. D 56,


24332434 1997.
17
D. Colloday and A. Kostelecky, CTP violation and the standard model,
Phys. Rev. D 55, 67606774 1997.
18
R. G. Winter, Neutrino oscillation kinematics, Lett. Nuovo Cimento
30, 101104 1981.
19
B. Kayser, On the quantum mechanics of neutrino oscillations, Phys.
Rev. D 24, 110116 1981.
20
C. Giunti, C. W. Kim, and U. W. Lee, When do neutrino really oscillate?: Quantum mechanics of neutrino oscillations, Phys. Rev. D 44,
36353640 1991.
21
Y. N. Srivastava, A. Widom, and E. Sassaroli, Charged lepton and neutrino oscillations, Eur. Phys. J. C 2, 769774 1998.
22
J. J. Sakurai, Advanced Quantum Mechanics AddisonWesley, Reading,
MA, 1967; F. Halzen and A. D. Martin, Quarks and Leptons Wiley, New
York, 1984.
23
W. Greiner and B. Muller, Gauge Theory of Weak Interactions SpringerVerlag, Berlin, 1993.
24
C. Giunti, C. K. Kim, and U. W. Lee, Comments on the weak states of
neutrinos, Phys. Rev. D 45, 24142420 1992.
25
M. Blasone and G. Vitiello, Quantum field theory of neutrino mixing,
Ann. Phys. N.Y. 244, 283311 1995; E. Alfinito, M. Blasone, A. Iorio,
and G. Vitiello, Squeezed neutrino oscillations in quantum field theory,
Phys. Lett. B 362, 9196 1995.
26
V. Alan Kostelecky, The Status of CTP, Talk presented at WEIN-98,
Santa Fe, New Mexico, June 1998; hep-ph/9810365.

ADJUSTING THE SPECTROMETER


I sometimes regret that much of our modern apparatus, even for students, has all the interesting
difficulties removed beforehand. If a student is going to work with a spectrometer, I think it is
highly desirable that he should go through the process of adjusting the collimator, the telescope,
and eye piece himself. It is desirable that he shall go through the process of getting the grating
lines parallel with the axis of rotation. It is desirable that he shall know how to set the axis of the
telescope perpendicular to the axis of rotation. Once the spectrometer is adjusted, all of the good
of the experiment has been utilized. I do not think that the student learns much in the last act of
measuring the wavelength of light.
W. F. G. Swann, The Teaching of Physics, Am. J. Phys. 193, 182187 1951.

875

Am. J. Phys., Vol. 67, No. 10, October 1999

Elisabetta Sassaroli

875

You might also like