Sei sulla pagina 1di 14

Introductory Nuclear Physics, K.

Krane
Conceptual Notes by Mordechai Rorvig
2009
1 General notes
Conservation of nucleons for
A
Z
X
The Q-value for a nuclear reaction is the change in rest mass energy, by convention,
Q = (m
initial
m
final
)c
2
.
Also, Q gives the new energy of the ssion products (including energy of particle emissions)?
(See homework 2, problem 5).
Q must be positive for any decay process. Best way to think about this is Q positive
corresponds to descending downwards a rung on one of the nuclear energy line diagrams.
Radiation can be alpha particles (helium nuclei, 2 neutrons and protons), beta particles
(electrons or positrons), gamma particles (high energy (frequency) photons) and neutrons, as
well as lots of dierent other kinds of particles with smaller probabilities.
Minimum Z expression for decaying nuclides based on liquid drop model.
In nuclear decay, the total number of nucleons stays the same, but the number of protons
change as they are converted to neutrons or vice versa. There is also orbital electron capture.
Unlikely that electrons are in nucleus, becuase they would need to be bound to protons by
a stronger force than the strong force. Uncertainty would require them to have extremely high
energies, but these are not observed. Total intrinsic angular momentum implies no electrons.
Magnetic moment of nucleus is observed to be much smaller than electron magnetic moment,
implying no electrons.
Isotopes are nuclides with same proton number, but dierent neutron number. Isotones are
the converse, nuclides with the same neutron number but dierent proton number. Isobars are
nuclides with the same mass number.
The syntax (A,B) denotes a reaction where a particle A enters a nucleus and a particle B
leaves the nucleus.
1.1 Relativistic dynamics
For relativistic dynamics p ,= mv and instead
p =
mv

1 v
2
/c
2
and also
T = E mc
2
where E =

(pc)
2
+ (mc
2
)
2
. For a photon, we have simply that
E = pc.
1
2 Chapter 2: Elements of quantum mechanics
2.1 Quantum theory of angular momentum
The angular momentum is a constant of motion given by L
2
= l
2
) = l(l + 1)
2
, for a wave
function given by R(r)Y
lm
l
(, ). The atomic substates with a given l are labelled s, p, d, f, g, h, i
for l = 0, 1, 2, 3, 4, 5, 6. By convention we usually choose the z component as determined, and
compute l
z
= m
l
. Note that m
l
= 0, 1, 2, . . . , l. Note that l
z
< l strictly (uncertainty
principle).
We also have the fundamental intrinsic angular momentum, or spin. We have S
2
= s(s+1)
2
and s
z
= m
s
, where m
s
= 1/2.
The total angular momentum is given by J = L+S. We have J
2
= j(j +1)
2
and j
z
= m
j
,
where m
j
= j, j +1, . . . , j 1, j. Also, m
j
= m
l
+m
s
. The J value is usually indicated with
subscript in spectroscopic notation. The quantum number n is indicated as a coecient, such
as 3p
3/2
.
The number of protons/neutrons that can go in each level is 2(2l + 1). The second term
from the m
l
degeneracy, the rst term from the m
s
degeneracy.
Odd parity if L is odd, even if L is even.
The magic numbers are 2, 8, 20, 28, 50, 82, 126, 184.
The angular momentum l has values l = 0, 1, 2, . . . , n 1.
A nucleus with odd A will have its ground state spin and parity detemined by the single
unpaired proton or neutron. A nucleus with even A with an unpaired neutron and proton will
have the angular momentum given by their absolute value of their dierence (the minimum
branch) and the spin given by the the multiplication of their parities (-+ odd, even, ++
even).
3 Chapter 3: Basic nuclear structure
3.1 The nuclear radius
3.2 Mass and abundance of nuclides
3.3 Nuclear binding energy
Binding energy is the dierence in mass energy between a nucleus
A
Z
X and its constituent Z
protons and N neutrons
B = Zm
p
+Nm
n
[m(
A
X) Zm
e
]c
2
= [Zm(
1
H) +Nm
n
m(
A
X)]c
2
.
Since B is generally positive, this is like saying the energy of the component parts is greater
than the energy of the combined nucleus. In other words, the combined nucleus is more stable
than the sum of the individual parts, so it has a lower energy.
The neutron separation enery is the amount of energy needed to remove a neutron from a
nucleus
A
Z
X. Hence
S
n
= B(
A
Z
X) B(
A1
Z
X) = [m(
A1
Z
X) m(
A
Z
X) +m
n
]c
2
.
Semi-empirical mass formula volume energy, surface energy, coulomb energy, and as-
symetry energy.
2
4 Chapter 4: Basic nuclear structure
5 Chapter 5: Nuclear models
Dicult to classically model the nuclear forces of heavier nuclei. Firstly, because the classical
models like the square well potential would require numerical solution. Secondly, because there
is evidence to show the interactions depend on three-body forces, for which there is no classical
analog. (Is this where QFT comes in?)
Thus we use models, which can be judged on how well they describe the measured phenom-
ena, and on how well they predict additional phenomena.
5.1 The shell model
In the atomic shell model (electron model), we ll up electrons in the outer shells according
to the Pauli principle. Then we obtain a core of lled shells and an exterior shell containing
valence electrons. We nd that the atomic properties vary smoothly for changing the valence
electrons within the subshell. But we nd strong changes when a subshell is lled and we begin
lling another. (For example, we can see this behavior in measuring the atomic radius vs. Z,
or the ionization energy vs. Z.)
In the atomic (electron) model, the potential is provided by the nuclear Coloumb force, so
we can solve the Schrodinger equation to calculate allowed energies and bound states. In the
nucleus, however, the potential is generated by the self-interaction, so we cannot do the same
thing.
Electrons have well known spatial orbits, regions in which collisions are rare. Nucleons are
much more conned.
We wind up seeing very similar behavior in the nuclear properties (e.g. separation energies
of proton and neutrons.) We see shells lled up at the magic numbers for Z equal to 8, 20,
28, 50, 82, and 126. We see even more stable congurations for doubly magic numbers when
both Z and N are magic numbers (e.g., 2 and 8).
This justies the usage of a shell model for the nucleus. Regarding the question of a nuclear
potential, we make the following fundamental assumption. There exists a nuclear potential for
a single nucleon caused by all the other nucleons. Then we can solve the Schrodinger equation
to provide information about energy levels.
The existence of denite spatial orbits is implied by the shell model. If two particles in a
lower subshell had a collision, they could not move up to the next subshell, because in general,
it would already be lled. Hence the nucleons would need to energize all the way to the valence
band, which would require more energy than is probable. Hence, the low (potential) energy
nucleons can be considered collisionless, with denite spatial orbits.
Using the 3-dimensional nite square well and harmonic oscillator potentials, we can solve
for the energy levels. We see the magic numbers 2, 8, and 20, but then they start deviating
from experiment. As an improvement, we can use an intermediate potential of the form V (r) =
V
0
1+exp[(rR)/a]
. This removes the l degeneracies of the major shells. R gives the mean radius
(R = 1.25A
1/3
) and a gives the skin thicknes a = .524 fm, the distance over which the potential
changes from .9V
0
to .1V
0
.
3
6 Chapter 6: Radioactive decay
Radioactivity motivated the birth of nuclear physics. Long nuclear decay times are responsible
for the presence of most radioactive elements in nature, such as uranium.
Radioactive isotopes can also be created by bombarding stable atoms with nuclear particles.
6.1 The radioactive decay law
Radioactivity discovered in 1896. Then it was discovered it followed an exponential law. Also,
it was discovered that changes took place to individual atoms, and not sample as a whole.
Finally, it was also discovered to be a statistical process with some degree of randomness.
The decay or disintegration constant
=

N
N
is the probability per unit time for the decay of an atom. The basic assumption of the theory
of radioactive decay is that this value remains constant (for each atomic conguration) for all
time. Note that 0 for stable nuclei.
Here we have several basic denitions, derivations and equations (see pg. 162-163).
The half-life of a sample is the time for half of the atoms to decay
t
1/2
=
0.693

.
The activity of a sample is the rate at which decay occurs in the sample
A(t) = N(t).
Note that the activity as derived requires that t t
1/2
.
The activity (in decays per second) diminishes exponentially so it can be used to track the
amount of sample and measure the half-life of the sample. However, the halife must not be
too long or too short.
A sample may decay in two or more ways; then it has partial decay constants,
a
+
b
=
with the total decay rate given by

N =

N
a
+

N
b
= N
a
+ N
b
= N. Note that we never
observe the partial decays; we only observe the total decay rate and total decay constant.
6.2 Production and decay of radioactivity
Assume we have a target of stable nuclei placed into an accelerator or reactor. Then some
atoms will undergo excitation to a radioactive state. This rate is given by
R = N
0
I
, depending on the original number of atoms, the current or ux of incident particles, and the
reactor cross section , which represents the probability of an incident particle to react with a
target nucleus. A typical ux might be 10
14
/scm
2
and a typical cross section might be a few
barns (10
24
cm
2
). Hence the probability to convert a particle is very small (10
10
) on typical
time scales, and we can consider N
0
constant.
Thus, for the new radioactive nuclei, the decay equation includes a formation term due to
R, and a new decay term corresponding to the decay constant. We have
dN
1
= Rdt
1
N
1
dt,
4
N
1
(t) =
R

1
(1 e

1
t
).
For small times, one can show the activity increases at a linear rate (expanding the ex-
ponential term.) For longer times, the activity approaches a constant, an example of secular
equilibrium.
Further irradiation thus provides diminishing returns as we reach the equilibrium value
where the rate of decay equals the rate of radioactive formation. After 3 halives we have 87.5%
irradiation.
6.3 Growth of daughter activities
Here we consider the mechanics of the reaction rates for a parent, daughter, and granddaughter
product, and various approximations we can make on the equations. Basically we confront a
basic uncoupled system of rst order ODEs with real exponential solutions (exponential decay
solutions).
If we generalize to n reaction products, we are left with the Bateman equations, which are
just a system of n rst-order ODEs, with coecients given neatly by a neat product rule.
6.4 Types of decays
The most common types of decays are , , and decays. In and decays the nucleus
emits a particle in order to decay to the most stable isobar for the resulting mass number. In
decays, a particle is emitted in order to decay down from an excited state, but leaving the
nuclear species unchanged.
Because the nucleus is such a tightly bound system, its binding energy is maximized, so
an emission maximizes the energy release.
In decay, either a proton is converted to an neutron (creating and emitting positron), a
neutron is converted to a proton (creating and emitting an electron), or an electron is captured
by the nucleus (electron capture) which converts a neutron to a proton. In all processes, a
chargeless neutrino is also emitted.
In emission an excited state decays to a lower excited state (or ground state). The half
lives are quite short, generally less than 10
9
seconds, but sometimes they can me buch longer,
(hours or days), and these are known as isomeric transitions (with corresponding isomeric
states).
The process of internal conversion competes with gamma emission, where excitation energy
is transferred to the ejection of an atomic electron, ionizing the atom. More specically, the
electromagnetic elds of the nucleus interact with the electron shell, ejecting an electron, when
the excitation energy is greater than the binding energy. This causes x-rays, auger electrons,
and competes against gamma decay.
Spontaneous ssion can occur with nuclear fragments distributed over the entire range of
medium weight nuclei.
As we look at heavier and more unstable nuclei, it can become more energetically favorable
to decay by nucleon emission. This occurs most frequently for ssion products which can have
a large excess of neutrons and are responsible for the delayed neutrons which are used to control
a nuclear reactor.
In reality, an excited nucleus has many dierent competing decay modes. The branching
ratios are given by the partial decay constants or partial half-lives, as discussed in section 6.1.
5
6.5 Natural radioactivity
Due to radioactive decay, we only observe naturally occuring radioactive elements on earth
which have half-lives that are long compared to the lifetime of the earth.
Since an emission involves a change in A by 4, we can classify radioactive decay chains in
terms of 4n,4n + 1,4n + 2,and 4n + 3 numbers.
Other sources of natural radioactivity include the atmospheric elements formed by collision
with cosmic rays (high energy protons).
6.6 Radioactive dating
The radioactive decay process involves samples with huge numbers of particles, i.e., N 1.
If we ask how many particles decay? in a timescale comparable to one half-life, then the
uncertainty in the binomial distribution goes like

N, while the standard . . .
7 Chapter 7: Detecting nuclear radiations
The basic principle of operation for a detector is radiation enters a material, interacts with
the native nuclei, releases its typically low energy atomic electrons and then these are used to
produce a voltage or current in the detector. The choice of material depends on the kind of
radiation we want to detect. For alpha particles, 1-100 m thickness; electrons, .1 to 1 mm; for
gamma rays, even 5 cm may not be enough.
A geiger counter can show the radiation is present, but to detect where it is, we need a
detector sensitive to position; to measure the energy, we need an output pulse proportional to
the energy of the radiation, and so on. There are all sorts of dierent requirements/options for
dierent scenarios for dierent detectors. No one detector can do them all.
7.1 Interaction of radiation with matter
7.1.1 Heavy charged particles
Coulomb scattering of charged particles by nuclei (Rutherford scattering) is not as important
for interactions with bulk matter. This is because the nuclei of the atom occupy only about
10
15
of the volume of their atoms. Hence most interaction is between charged particles and
atomic electrons. Since the loss in kinetic energy for a direct collision is only
T = T
4m
M
(M the heavy ion mass, m the electron mass) it takes many thousands of collisions before a
charged particle of several MeV loses all its kinetic energy.
Furthermore, in a glancing collision, the charged particle almost follows a straight path.
Thirdly, since the Coulomb force has innite range, the particle loses energy continuously along
its entire path, and hence has a denitive range of penetration in the material. This range
varies little so that the mean range is fairly well dened for the same particle, material, and
energy. Fourthly, the particle can often ionize the atomic particles, causing possibly secondary
radiations (delta rays) which cause more ionizations.
The range scales with density and atomic mass number, and there is also a quantum me-
chanical formula for the stopping power dE/dx of a bulk material. However, this formula fails
at low energy as it does not consider the capture of electrons by the slow moving particle.
6
7.1.2 Electrons
Electrons have some key dierences from heavy charged particles. First of all, since they are
so much lighter, they have large angular deections, and so they do not follow a well dened
(straight) path in the material, and their range must be empirically determined.
Secondly, their low mass means they are subject to high accelerations, and for high energies
they can undergo substantial radiation damping, besides just coulomb damping (collisional
energy losses). These losses are only signicant at high energies and in heavy materials.
7.1.3 Electromagnetic radiation
Gamma rays and x-rays interact with matter through photoelectric absorption, compton scat-
tering, and pair production.
In photoelectric absorption, the photon is absorbed by an an atom and one of the atomic
electrons (photoelectron) is released. This cannot occur for a free electron as conservation of
energy requires the presence of a heavy atom to absorb the momentum lost. The kinetic energy
is T
e
= E

B
e
. This eect is most signicant for low energy photons ( 100 keV), increases
rapidly with Z ( Z
4
), and decreases rapidly with increasing photon energy ( E
3

). Furthermore,
the cross section (probability for absorption) has discontinuous jumps at binding energies of
electronic shells. Once the energy increases just past a shell binding energy, there is suddenly
a much higher availability for those shell electrons to take part (K edge).
In compton scattering, a photon scatters from a nuclear electron, resulting in a scattered
electron and a less energetic photon. If we regard the stuck electron as free and at rest (good
approx.), we can use relativistic conservation to nd a formula, for
E

=
E

1 + (E

/mc
2
)(1 cos )
.
Probability can also be given by QM calculation (Klein-Nishina formula.)
In pair production, a photon creates an electron-positron pair; the energy balance is E

=
T
+
+ mc
2
+ T

+ mc
2
. Obviously there is a threshold energy 2mc
2
for pair production so
this only tends to be dominant for higher energies. Like photoelectric absorption, it requires a
nearby heavy nucleus to absorb the momentum loss.
If we consider a beam of photons on a slab of thickness t, we have as a total linear
attenuation coecient, where simply = + + (for photoelectric absorption, compton
scattering, and pair production losses, respectively). The fractional loss in intensity is
dI
I
= dx
so that
I = I
0
e
t
.
7.1.4 Conclusion
In summary, consider looking at the losses of a monoenergetic 1 MeV beam of alphas, electrons,
and gamma radiations. The alpha intensity drops sharply at the mean range length (.0003 cm);
the electron insensity decreases slowly even for thickness much less than the range ( 0.18 cm
extrapolated total). And the gamma intensity decreases exponentially, but penetrates much
further (4.3 cm).
7
8 Chapter 8: Alpha decay
Rutherford discovered particles and showed that they were actually helium nuclei.
Only extremely rarely are other nucleons besides alpha particles (such as a deuteron nuclei)
emitted from the nucleus. There is a special reason why alphas tend to be emitted (see below.)
8.1 Why alpha decay occurs
The alpha particle is especially tightly bound and stable. If we look at a table of Q values for
various nuclide emissions, most of the values are negative. For emission, Q > 0, meaning
energy is released daughter nuclide more stable.
There are several other nuclide emissions with positive Q, but these are for heavier nuclides
(e.g.
8
Be) which have very low partial disintegration constants. Hence these emissions are never
seen. Conversely, sometimes emission can overwhelm the emission.
For heavier nuclei, A > 190 most are unstable against emission but only about one-half
meet the decay constant requirement.
8.2 Basic decay processes
Alpha decays typically release about 5 MeV energy. Thus T mc
2
and we can safely use
non-relativistic analysis. Using conservation of energy (for a parent nuclide initially at rest)
and momentum, we get
(m
X
m
x
m

)c
2
= T
X
+T

= Q.
(The decay will occur spontaneously only if Q > 0.) We can also nd T

= Q/(1 + m

/m
X
)
and using A 4, T

= Q(1 4/A).
The recoil energy of the daughter nuclide is generally only about 2% of the Q value, but this
can still be signicant enough to send it out of the original material. (This can be inhibited by
putting a surface coating on the radioactive material).
Note that the energy of the alpha particle can be measure, allowing us to nd m
X
, (making
some assumption like T

= .98Q, I suppose.)
8.3 Alpha decay systematics
There is an extremely strong correlation between increasing alpha energy and decreasing halife.
A factor of 2 in energy leads to a factor of 10
24
in half-life! This is the Geiger-Nuttall rule.
8.4 Theory of alpha emission
More mathematical section. See pg. 253. We consider the alpha to be preformed inside the
nucleus. We consider a decay constant for the alpha particle by setting = fP, where f is the
frequency the alpha particle touching a nite well barrier, and P is the probability for tunneling
through it. We can nd P by looking at an energy balance and a transmission coecient, and
we can nd f by looking at an average velocity and distance.
We neglect angular momentum and we assume the nucleus is spherical. The fact that the
nucleus is non-spherical for higher A actually leads to signicant error, since the radius can no
longer be simply approximated.
If we look at the emission for
14
C, its decay constant is less by only 10
3
, but its observed
decay is less by 10
9
. Thus it is 10
6
less likely to be preformed in the atomic nucleus.
8
Proton emission is usually prohibited because Q < 0. However, for special cases like Z A,
proton emission is possible.
8.5 Angular momentum and parity in alpha decay
I = [I
i
I
f
[, and if there is change in parity, this is given by (1)
l
, so it implies l must be
odd, e.g., for 0 I 4, 2
+
2

, we only only allow l = 1 or 3.


9 Chapter 9: Beta decay
Three processes are grouped together under this title. One, where a neutron is converted to
a proton, and an electron is created inside the nucleus and immediately ejected (

). The
converse process;
+
. Also, electron capture, where an atomic electron is captured by the
nucleus and a proton is converted to a neutron. A neutrino or antineutrino particle is also
involved in each decay.
9.1 Energy release in beta decay
There is a continuous spectrum of energy of emitted particles in beta decay, unlike alpha decay,
where you have a unique alpha energy corresponding to every combination of initial and nal
nuclear states. A rst suggestion that this was due to interaction with atomic electrons (on
ejection) was experimentally ruled out by calorimetric experiments. The eect was found to be
due to the emission of the neutrino (antineutrino), a very low mass particle that steals some
of the emitted electron energy.
An energy analysis on a free neutron decay shows that the neutrino can be considered
essentially massless. This follows on measurements of the maximum kinetic energy of the
emitted electron. Related to this, the neutrino moves at the speed of light and suers negligible
particle interactions. Hence the neutrino must be treated relativistically.
For
+
decay we wind up getting Q subtracts 2m
e
. But nor for

or electron cature.
Experimental momentum measurements of the resulting particle decays conrms that a
third particle must be present in order for conservation of momentum to be satised.
We can thus work out Q formulas for each type of beta decay. In these formulas, we start
out by working with the nuclear masses instead of the atomic masses. For electron capture,
we must furthermore take into account the reduction in atomic binding energy of the daughter
nucleus due to the loss of a shell electron. Also, if we decay to excited states, the Q value is
reduced accordingly.
9.2 Fermi theory of beta decay
9.3 Angular momentum and parity selection rules
Allowed decays. We regard the electron and neutrino wave functions as originating from the
origin, so they carry no orbital angular momentum. In other words, we have l = 0. Thus the
only change in the angular momentum of the nucleus results from the spins, each of which has
value s = 1/2. If they are antiparallel (S = 0), this is known as Fermi decay, then for l = 0
there can be no change in nuclear spin (I = [I
i
I
f
[ = 0). If they are parallel (S = 1),
this is known as Gamow-Teller decay, and thus I = 0, 1. In GT decay, at least one unit of
angular momentum is carried away (or added in). Since the electron and neutrino states carry
9
no angular momentum, the emitted l value is zero relative to the nucleus, so the parity (given
by (1)
l
must be even (no parity change.)
We therefore have the selection rules for allowed beta decay; I = 0, 1 and =no.
See examples on pg. 289, where we have some F transitions, some GT transitions, and
some mixed F/GT transitions which satisfy both rules. For example, n p has I = 0,
(1/2+ 1/2+).
Forbidden decays. These refer to the less likely (but still possible) cases in which l ,= 0.
The most frequent occurence is when the initial and nal states have dierent parities, so the
conditions for allowed decays are violated. Then, in rst forbidden decays, we take that one
unit of angular momentum is carried out. Thus Fermi decay types carry one change in angular
momentum, while GT decay types carry I = 0, 1, 2. And similarily for second forbidden
decays and so on.
We therefore have the selection rules for rst forbidden beta decay; I = 0, 1, 2 and =yes.
For the second (and later) forbidden decays, we only consider cases where I = 2, 3, as if it
equaled to 0 or 1, the selection rules are satised for the allowed decays, which are much more
probable.
9.4 Beta-delayed nucleon emission
After beta emission or electron capture, it is sometimes energetically favorable to emit a nucleon
as a decay mode. Thus we have a decay sequence called alternatively, parent to emitter to
daughter, where this process can occur. More specically, this decay sequence occurs when
the beta decay leads to an excited state of the emitter. When this excited state is higher than
the nucleon separation energy (for emitter to daughter), there is enough free energy for nucleon
emission.
In general this can have the eect of increasing the observed half-life for these nucleon
emissions (hence the name).
10 Gamma Decay
Most alpha and beta decays and nuclear reactions leave the nal nucleus in an excited state,
which then decays by gamma emission. Gamma rays are high-frequency photons (e.g., 10
19
).
Their wavelengths are very short (e.g., 10
4
and 100 fm) much shorter than say, visible light.
They have energies between 0.1 and 10 MeV, corresponding to the dierences between nuclear
energy levels.
Since they have negligible absorption and scattering in air, they are relatively easy to ob-
serve, and so they provide one of the most common methods for nuclear spectroscopy.
10.1 Energetics of gamma decay
Performing an energy analysis on a gamma decay process give an expression of the (relativistic)
energy of the gamma particle. Comparing this expression with known energy dierences (E =
E
i
E
f
) reveals that E

E. The exception occurs for very high energy gammas (5-10 MeV)
for which the recoil correction is signicant and can actually cause a displacement of the atom
in the solid lattice.
10
10.2 Classical electromagnetic radiation
The nuclear states can be modeled based on their multipole expansions. Each multipole gives
a dierent radiation distribution in space, so that by measuring the radiation distribution, we
can identify which moments are dominant. By considering the multipole expansions, we nd
an average radiated power per unit time for electric and magnetic poles.
We dene the index of the radiation by L where 2
L
is the multipole order (L = 2 gives
quadrupole order, etc.) and we denote the states as ML and EL, e.g., M1,E3, and so on. The
angular distributions for the L states are given by the legendre polynomial P
2L
(cos ). The
parity of the radiation eld is given by (ML) = (1)
L+1
and (EL) = (1)
L
. The radiated
power is given by an expression on page 331.
10.3 Transition to quantum mechanics
Here we nd the Weisskopf estimates for the decay constants for each multipole. These show
that the lower multipolarities are dominant and the electric multipolarities are stronger for
higher A.
10.4 Angular momentum and parity selection rules
Here we have [I
i
I
f
[ L I
i
+ I
f
and if is yes or no, we must choose l values such that
the M and E elds are correspondingly even or odd. (e.g., if parity change is no, then we must
have only M1, M3, and E2, E4, etc. etc.
11 Chapter 11: Nuclear reactions
When nuclear particles collide with bulk material, they can interact to cause nuclear reactions.
These were rst observed by Rutherford (like everything...) with alpha particle collisions with
nitrogen. The emitted particles can come from accelerators, reactors, or even from radioactive
decays. In this book we focus on low energy reactions on the order of 10 MeV per nucleon
or less. Medium energy and high energy reactions can result in exotic particle transitions (e.g.,
neutrons convert to protons) and particle emissions (e.g. muons), and even rearrange quarks
that constitute all nucleons.
11.1 Types of reactions and conservation laws
Note we represent reactions like X + a Y + b as X(a, b)Y , where X is the target, a is the
accelerated projectile, and Y and b are the reaction products.
Scattering reactions are those where the incoming and outgoing particles are the same (as
are X and Y). Elastic processes are those where Y and b are in the ground state, and inelastic
processes are those where Y or b is in an excited state (then typically decay quickly by gamma
emission). When a and b are the same particle, but another nucleon is ejected separately, then
it is called a knockout reaction (e.g., compton scattering.) Transfer reactions are those where
one or two nucleons are transferred between projectile and target.
In direct reactions, only one or two nucleons take part in the process. In compound nucleus
mechanism, the incoming and target nuclei merge for a complete sharing of energy before b
is ejected. In between are resonance reactions where the incoming nuclei forms a quasibound
state before b is ejected.
11
11.1.1 Observables
Energy (resolution 10 KeV with magnetic spectrometer), angular distribution, dierential cross
section (probability to observe particle b with a certain energy at a certain angle), total cross
section (probability to be emitted at certain energy), absolute total cross section (probability
for Y to be produced).
Polarization experiments to deduce spin, observe gamma radiations or conversion electrons
from Y and also b, can observe angular distribution of these gammas to understand their
spin-parity states.
11.1.2 Conservation laws
Conservation of total energy and linear momentum, conservation of proton and neutron number
(due to the low energy at higher energies, only total nucleon number conserved). Conservation
of angular momentum and conservation of parity (total parity before equal to total parity after.)
11.2 Energetics of nuclear reactions
We dene the conservation of energy as usual, and Q as usual. If Q > 0 we call the reaction
exothermic or exoergic and energy is released. If Q < 0 then kinetic energy is converted into
mass or binding energy of the reaction products.
Assuming T
Y
= 0, we can get a formula for T
b
in terms of . We can plot T
a
vs. T
b
and
we get linear correlations for dierent angles. Except at low energies, where Q < 0, we have a
threshold value of T
a
for which the reaction will not occur, which is always when = 0. There
is also a double valued situation for Q < 0 which is only important for reactions involving nuclei
of comparable masses.
If the reaction reaches an excited state, we must put E
ex
on the RHS of the energy conser-
vation equation and modify Q accordingly.
12 Chapter 12: Neutron physics
Not aected by Coulomb barrier, so they can easily penetrate the nucleus and cause nuclear
reactions. Dicult to measure; only negligibly interact with electrons, so they dont ionize
atomic electrons and cause other radiation to be emitted. Free neutrons are unstable against
beta decay, with a halife of 10.6 minutes.
12.1 Neutron sources
We cant accelerate neutrons directly, but we can allow them to intercept other materials
(moderation) and slow them down to selected energies. Classify neutrons as 1. Thermal (.025
eV), 2. Epithermal (1 ev), 3. Slow (1 keV), 4. Fast (100 keV - 100 MeV).
One source is the alpha-beryllium source. Beryllium (A = 9) has a relatively loosely bound
neutron, and if a typical alpha particle (5-6 MeV) strikes a beryllium nucleus, a neutron can be
released. If we mix together a long lived radioactive element which emits alphas (e.g.,
226
Ra)
with beryllium, we have a steady neutron source. These neutrons are not monoenergetic due
to the dierences in collision angles and the varying moderation of the alpha particle energies
by material collisions. Also, the daughter nucleus can be left in an excited state.
12
Similarly, we can also use photoneutron sources where gamma emitting isotopes are com-
bined with an element like beryllium. Since gamma emissions have discrete energy levels, this
can produce a more monoenergetic neutron ux.
Thirdly, neutrons can be produced from the spontaneous ssioning of isotopes (e.g.,
252
Cf.
These neutrons have a continuous energy spectrum characteristic of ssion, ranging from 1 to
3 MeV. Fourthly, we can use nuclear reactions, such as by those produced from accelerator
collisions. By selecting the emission angle and the incident energy, we can choose a monoener-
getic beam of almost any energy. Fifthly, we can also use nuclear reactors, which have a large
neutron ux of around 10
14
neutrons/cm
2
/s. Most of the neutrons have been moderated to
thermal energies inside the reactor, but some fast neutrons are also emitted. These neutrons
have an energy ranging from about 1 to 7 MeV, peaking at 1-2 MeV (characteristic of ssion
emission intensities.)
12.2 Absorption and moderation of neutrons
12.3 Neutron reactions and cross sections
13 Chapter 13: Nuclear ssion
In 1939 Hahn and Strassman showed that intermediate mass nuclei were formed in the bom-
bardment of uranium by nucleons, and showed that the released energy was high, of the order
of 100 MeV.
Fission results due to competition between nuclear binding force and Coulomb force; binding
force scales with A, but Coulomb force scales with Z
2
, so at higher atomic numbers, heavy nuclei
have a relatively smaller potential to overcome. (Nucleus feels a stronger internal repulsion).
Can occur spontaneously as a natural decay, or due to the absorption of a ligher particle
such as a neutron or photon.
The chain reaction eect can allow for a large total energy release, because each ssion
sends out neutrons which can incur the ssion of other atoms, either rapidly (e.g., in a bomb)
or slowly (e.g., in a reactor).
13.1 Why Nuclei Fission
A heavy nucleus dividing into lighter nuclides increases the binding energy per nucleon (gure
3.16.) (Since iron has the highest binding energy per nucleon.) Thus the new nuclides are more
tightly packed. This process releases energy, i.e., it is preferential, allowing the nucleon to reach
a lower energy state.
However, decay is inhibited and not preferred for Z < 250 because the Coulomb barrier
(force of repulsion between two nuclei) to form the heavy nucleus (e.g., 250 MeV for U-238) is
higher than the amount of free energy that can be released in a spontaneous decay (e.g., 214
MeV).
Height of Coulomb barrier is roughly equal to energy released in ssion of heavy nuclei.
Spontaneously ssioning nuclei are nuclei where ssion competes successfuly with other
types of radioactive decay.
Calculations suggest ssion occurs automatically for A > 300.
For some nuclei, energy of thermal neutrons may be able to push them over the barrier, for
others, they may require fast neutrons.
Activation energy is the height of the ssion barrier above the ground state (i.e., energy
needed to overcome the barrier.) (Figure 13.3) (Table 13.1) This is the energy needed to achieve
13
ssion; if the excitation energy of adding a neutron is greater than the activation energy, ssion
can occur.
Liquid drop model can be used to show that stretching the nucleus can increase the total
energy for certain combinations of A and Z, and provides a criterion for ssioning.
13.2 Characteristics of Fission
Fission products are not determined uniquely determined.
For thermal neutrons, there is a distribution of masses, symmetric about its center, favoring
one heavy fragment and one light fragment. This is characteristic of low energy ssion. High
energy ssion favors fragments of equal mass.
Fission fragments shed excess neutrons at the instant of ssion (within 10e16 sec), known
as prompt neutrons. The average number of prompt neutrons is called . This mimics the
statistics of an evaporation process, and it is independent of the ssioning nucleus.
Delayed neutrons are emitted following decay of ssion fragments. Delay times are on the
order of seconds. Total intensity amounts to 1 neutron in 100 ssions.
The initial ssion products are highly radioactive and emit many and radiations until
decaying to stable isobars. For example, 93-Rb to 93-Sr to 93-Y to 93-Zr to 93-Nb.
13.3 Energy in Fission
Excitation energy is the energy gained when a nucleus captures a neutron to form a compound
state; for example, when U-235 captures a neutron to form U-236*, we have E
ex
= [m(
236
U

)
m(
236
U)]c
2
.
If we assume the captured neutron enters with a low energy (e.g., thermal neutron), then
we can neglect its kinetic energy altogether, and then the excited mass of the U-236* atom is
simply m(
236
U

) = m(
235
U) +m
n
.
If the excitation energy is less than the activation energy E
A
(the energy needed to excite an
atom into a ssionable state), clearly a thermal neutron will not be enough to induce a ssion.
14

Potrebbero piacerti anche