0% found this document useful (0 votes)
20 views

1FNP_Nuclear_ParticlePhysics

Uploaded by

Luis Perales
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
20 views

1FNP_Nuclear_ParticlePhysics

Uploaded by

Luis Perales
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 27

FNP

Nuclear & Particle Physics

Universidad de Córdoba

Marta León Miranda

2022/2023

Page: 1
FNP

Contents
1 Properties of the atomic nuclei 3
1.1 Introduction to Nuclear Physics . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Nuclear Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Shape factor of nuclear charge . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Nuclear mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4.1 Atomic mass units (u) . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4.2 Nuclear mass measurements . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Energy of the nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5.1 Binding energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5.2 The stability valley . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.5.3 Nucleon separation energy . . . . . . . . . . . . . . . . . . . . . . . 10
1.5.4 Semiempirical mass formula . . . . . . . . . . . . . . . . . . . . . . 11
1.5.5 Mass parabola . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.6 Nuclear Quantum Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.6.1 Wave function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.6.2 Nuclear spin and parity . . . . . . . . . . . . . . . . . . . . . . . . 13
1.6.3 Nuclear isospin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.7 Nuclear electromagnetic moments: dipole magnetic moment . . . . . . . . 14

2 Radioactivity and radiological protection 15


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Types of disintegration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Radioactive disintegration law . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Graphical representation of the mass parabola . . . . . . . . . . . . . . . . 16
2.5 Dosimetry and radiological protection . . . . . . . . . . . . . . . . . . . . . 17
2.6 Protection measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3 Nuclear models and nuclear phenomenology 19


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2 Macroscopic collective terms . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.1 Liquid drop model . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.2 Fermi gas model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.3 Even-even nuclei collective properties . . . . . . . . . . . . . . . . . 20
3.2.4 Vibrational model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2.5 Rotational model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3 Microscopic or single particle models . . . . . . . . . . . . . . . . . . . . . 21
3.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3.2 Extreme single particle model . . . . . . . . . . . . . . . . . . . . . 21
3.3.3 Shell model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

4 The nucleon-nucleon interaction 24


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2 Deuteron properties and wave function . . . . . . . . . . . . . . . . . . . . 24
4.3 Yukawa potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.4 Nucleon-nucleon potential . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

Page: 2
FNP

1 Properties of the atomic nuclei


1.1 Introduction to Nuclear Physics
Nuclear Physics is the scientific field which studies the atomic nuclei, their properties and
the forces that act among their constituents, protons and neutrons, commonly known as
“nucleons”. As the nuclei are physical entities unusually small (between 2 and 10 fm), their
study must be addressed employing tools of the Quantum Mechanics, although sometimes
one must resort to macroscopic concepts.

The nucleus is a complex system, formed by a large number of constituents in which


three of the four fundamental forces in Nature act: the nuclear force, also called “strong”
force because it is the most intense of the known interactions; the electromagnetic force
and the weak force.
Despite the great development of the Nuclear Physics in the 20th century, there is not
yet an available and acceptable theory that describes the properties of the nuclei and we
only have a phenomenological view. It is not possible to describe Nuclear Physics from
first principles as in Classical Mechanics or Electromagnetism.

Atomic nuclei are composed by A nucleons (also called “mass number”), with Z the
number of protons and N = A − Z the number of neutrons. Nuclei with values of A
ranging from 1 to 260 approximately can be found. Therefore, they consist of a large
number of nucleons that are usually called “nuclide”. The following notation is employed
to distinguish them:

ZX
A

For instance: 11 H, 21 H, 31 H, 235


92 U, 92 U.
238

There are 274 natural nuclides, all of them stable and present on Earth, and more than
2500 unstable nuclei have been studied (also called “artificials”). Nuclei with the same
atomic number Z are named “isotopes”; with the same number N , “isotones”; and with the
same number A, “isobars”. The only isotopes that have been named with a different name
from their chemical element have been those associated to Hydrogen, called “deuterium”
(A = 2) and “tritium” (A = 3).
The corresponding nuclei are called proton (p), deuteron (d) and triton (t). The nucleus
of the Helium 4-atom ( 42 He) is identified as the alpha particle: α.

Free nucleons have the following properties:

mass mean life charge


proton mp = 1.673 × 10−27
kg = τp > 5 × 10 years qp = +e
32

= 938.3 MeV/c2
neutron mn = 1.675 × 10−27 kg = τn = 885.7 ± 0.85 qn = 0
= 939.6 MeV/c 2

Table 1: Properties of the free nucleons

DUDA: ¿FALTAN UDS MEAN LIFE NEUTRON?

Page: 3
FNP

Neutron, slightly more massive than proton, is unstable when it is free. On the
contrary, no case of proton disintegration has been found yet, reason why its mean life
(τp ) is usually given as an upper limit.
Nucleons have around 1840 times more mass than electrons, so, the total mass of the
atom consists almost entirely of the nucleus. The employed charge is e = 1, 6 × 10−19
C (absolute value of the electron charge). Nucleons are fermions, i.e., they have a half-
integer spin, in particular, s = 1/2. Their magnetic dipolar moments are µp = 2, 79µN
and µn = −1, 91µN , with:

µN = eℏ
2mp
= 3, 152 × 10−14 MeV/T

the nuclear magneton.


The isospin of the nucleon is t = 1/2. By agreement, proton is the state with the third
component t3 (p) = +1/2 and neutron, t3 (n) = −1/2.

The stability of the nuclei implies the existence of a Nuclear Force, which is attractive
and binds all the nucleons, it is more intense than the electrostatic repulsion and of short
range (of the order of 1fm approximately). Nuclei are considered quantum systems with
well defined static and dynamic properties:

• Mass M , radius R and atomic number Z (which determines its electric charge).

• Spin J and parity P , usually expressed as J P .

• Isospin T .

• Dipolar magnetic momentum µ and electric quadrupole Q.

• Mean life τ .

• Scattering cross section σ.

• Discrete spectrum of energy levels En .

Among the electromagnetic multipole moments of the previous list, the electric dipolar
moment does not appear because it is zero, due to the charge symmetry.

In normal conditions, a nucleus will be in its ground state, which is the minimum state
energy and equal to the nuclear mass. This is the most accessible state and, therefore,
the easiest to study.
It can be checked that in any nuclear reaction, A, Z and N are conserved.

1.2 Nuclear Structure


The nucleus is roughly spherical. Near to the surface, the outline is imprecise and the
density decreases gradually. It is what is known as “nuclear skin”. The charge of matter
distribution is characterised by two parameters: the radius (R) and the skin thickness (t).

The obtained experimental results when measuring the charge nuclear density lead to
the Fermi or Saxon-Woods distribution:

Page: 4
FNP

ρ0
ρ(r) = r−R (1)
1+e t
with ρ0 the maximum nuclear charge density and

R = r0 A1/3 (2)

where r0 denotes the mean “size” of the proton in a nucleus (remember that for Hydrogen
A = 1) and its value is r0 ≃ 1, 22 fm; t is related to the size of the nuclear skin, it is
essentially the same for all nuclei and its is t = 0, 55 ± 0, 07 fm. The radius R is the value
in which the density is reduced to its half, ρ(R) = ρ/2. The Saxon-Woods distribution is
displayed in figure (1).

Figure 1: Saxon-Woods distribution for R = 5fm and t = 0, 5fm.

1.3 Shape factor of nuclear charge


The differential cross section dΩ

(being Ω the solid angle) is a key magnitude in scattering
processes. When the target of the collision  does not have spatial extent, the point-like
differential cross section is considered, dΩ p .

However, in the case in which the target has a spatial extent, this previously described
section must be multiplied by a function which takes into account that extent. This
function is called “shape factor” and is usually represented by F (q 2 ), resulting:
 
dσ dσ
= × |F (q 2 )| (3)
dΩ dΩ p

The shape factor is defined as the Fourier transform of the nuclear charge distribution
ρch (⃗r): Z
1 ⃗·⃗
q r
2
F (q ) = ρch (⃗r)ei ℏ dV (4)
Ze R3
with ρch (⃗r) = Ze|ψ(⃗r)|2 and the ensuing normalization under spherical symmetry:
Z Z ∞
Ze = ρch (⃗r)dV = 4π ρch (r)r2 dr (5)
3
| R {z } | 0 {z }
General formula Spherical coordinates

Page: 5
FNP

In expressions (4) and (5), the integration limits extend over all the space, denoted as
“R3 ”. If the launched projectile is an electron, ⃗q = p⃗i − p⃗f ; with p⃗i and p⃗f the initial and
final momentum of the electron, respectively. Finally, Ze is the charge of the nucleus.
The expression (4) can be written is spherical coordinates as follows:

4πℏ ∞
Z  qr 
2
F (q ) = ρch (r)r sin dr (6)
Zeq 0 ℏ

In the case of quasi-elastic collisions q −→ 0 and the sin (x) function in (6) can be
expressed as a Taylor series:

4πℏ ∞
Z  
2 qr 1  qr 3
F (q ) = ρch (r)r − + ... dr ≃
Zeq 0 ℏ 3! ℏ
4πℏ ∞ 4πℏ ∞
Z Z
qr 1  qr 3
≃ ρch (r)r dr − ρch (r)r dr =
Zeq 0 ℏ Zeq 0 6 ℏ
Z ∞
q2 1
Z
1 2
= 4π ρch (r)r dr − 2 ρch (r)r2 dV =
Ze 6ℏ Ze 3
| 0 {z } | R
{z }
Ze ⟨r2 ⟩

1 q2
= Ze − 2 ⟨r2 ⟩
Ze 6ℏ
that is:
q2 2
F (q 2 ) ≃ 1 − ⟨r ⟩ (7)
6ℏ2
In the case of a perfectly elastic collision q = 0 and F (0) = 1, so that the differential
cross section of a point-like target is obtained. For a spherical nucleus of constant charge
ρch (r) = ρ0 , it follows that:
R 2
R 2
ρ r dV 3 r dV
Z
1 3 0
⟨r2 ⟩ = ρ0 r2 dV = RR = RR =
Ze R3 R3 0
ρ dV R3
dV
RR RR 4 R5 (8)
4π 0 r2 r2 dr 0
r dr 5 3 2
= RR = RR = R3 = R
4π 0 r2 dr r 2 dr
3
5
0

Therefore:
(9)
p
R = 1.29 ⟨r2 ⟩
Examples of measurements obtained with this method are: ⟨r2 ⟩( 40 Ca) = 3.5 fm and
p

⟨r2 ⟩ 206 Pb) = 5.5 fm.


p

The distribution of measurements of the nuclear radii ⟨r2 ⟩ when fitted to the mass
p

number A1/3 gives rise to a constant for A ≥ 50, obtaining A < 50:

(10)
p
⟨r2 ⟩ = 0.95A1/3

Page: 6
FNP

1.4 Nuclear mass


1.4.1 Atomic mass units (u)
From an experimental point of view, it is much easier to measure the mass of the atoms
rather than the mass of the nuclei. What is commonly measured is the atomic mass
of ionized atoms, because their analysis in electric and magnetic fields is simple. As a
consequence, in the nuclear tables the atomic masses are usually found.
Since 1961, the atomic mass unit (u) is defined as the mean mass of a nucleon 12 C, it is,
Ma ( 12 C)/12, with Ma ( 12 C) the atomic mass of the 12 C atom. To know the equivalence
in kilograms, the number of atoms in one mole must be taken into account and turns out
that:
12NA u = 0, 012 kg
(11)
⇒ u = 1, 67 × 10−27 kg

which lead to the definition of the equivalent energy Eu :

Eu = uc2 = 931, 5 MeV (12)

Therefore, the atomic mass unit is equal to:

u = 931, 5 MeV/c2 (13)

1.4.2 Nuclear mass measurements


The atomic mass is not the addition of the nuclear mass plus the electron masses. The
binding energy of the electron would be missing and thus, it is right to say that:

(nuclear mass) + (electron mass) = (atomic mass) + (electron binding energy) (14)

If the atom is defined as AZ X and we denote: Mnuc = nucleus mass, Ma = atomic mass
and Bi = binding energy of the i-th electron, expression (14) can be rewritten as:
Z
X Bi
Mnuc (Z, N ) + Zme = Ma (Z, N ) + (15)
i=1
c2

In the previous formula, the electron binding energy can be neglected because it is as much
of the order of 0, 1MeV. Nevertheless, it is more convenient to carry out the calculations
with the “mass defect”, which is defined as:

∆(Z, N ) = Ma (Z, N ) − Au (16)

As A is a dimensionless magnitude, it multiplied by u in equation (15).

In nuclear reactions, it is likely to happen a change of mass from reactants to products.


If we consider the following reaction: a + A −→ b + B, the heat of reaction Q measures
the available energy and it is defined as the difference between the mass of the initial state
and the mass of the final state, multiplied by c2 , this being:

Q = Minitial c2 − Mfinal c2 =
(17)
= ma c2 + mA c2 − mb c2 − mB c2

Page: 7
FNP

On the other hand, due to the energy conservation, it is given that:

Q = Tb + TB − Ta − TA (18)

with Tb , TB , Ta and TA the kinetic energy of b, B, a and A, respectively. Usually, the


target A is at rest and, therefore, TA = 0.

• If Q > 0 ⇒ exothermic reaction.

• If Q < 0 ⇒ endothermic reaction.

1.5 Energy of the nuclei


1.5.1 Binding energy
In Nature, the energy of a system made up by several components is less than the addition
of the energy of each component separately, normally, it is formed a “bound state”. If we
consider the nucleus as a bound state and take into account that energy and mass are
equivalent (not equal), it follows that the mass of the components separately is equal to
the nuclear mass plus the binding energy (both masses multiplied by c2 ):

(mass of the nuclear components) × c2 = (nuclear mass) × c2 + (binding energy) (19)

As the components of the nucleus are Z protons, each proton with mass Ma ( 1 H) and N
neutrons, each one with mass mn ; if we call “EL (Z, N )” to the nuclear binding energy,
expression (19) can be rewritten as:

ZMa ( 1 H)c2 + N mn c2 = Ma (Z, N )c2 + EL (Z, N ) (20)

in this equation, it is assumed that the nuclear mass is equal to the atomic mass.
Therefore, the binding energy EL (Z, N ) is:

EL (Z, N ) = ZMa ( 1 H) + N mn − Ma (Z, N ) c2 (21)


 

However, it is convenient to express the binding energy as a function of the mass defect
instead of the mass itself. To transform equation (21) we add and subtract Au in the right-
hand side of the equation and recall the definition of the mass defect given in equation
(16):
EL (Z, N ) = [Z∆H + N ∆n − ∆(Z, N )] c2 (22)
with:
∆H = Ma ( 1 H) − u = mp − u =
= 938.3 MeV/c2 − 931.5 MeV/c2 = 6.8 MeV/c2

∆n = mn − u =
= 939.6 MeV/c2 − 931.5 MeV/c2 = 8.1 MeV/c2
Instead of considering the binding energy EL (Z, N ), it can be studied the binding
energy per nucleon EL /A = ε, quantity displayed in figure (2).

Page: 8
FNP

Figure 2: Variation of the binding energy per nucleon, EL /A as a function of the mass
number A.

From this curve we find the following results:

• Its maximum value is 8, 7MeV for A = 56 (near the Iron) and drops down to 7, 6
MeV for A ≃ 240. The mean value is ε̄ = 8 MeV.

• As the number of nucleons increases, the binding energy per nucleon also does,
reaching a point in which it saturates and becomes nearly a constant value. This
fact reveals that it is a short-distance reaction.

1.5.2 The stability valley


The called “Segrè chart” (figure (3)), shows in the (Z, N ) plane all the studied nuclei,
both: stable or not. The shaded central area corresponds to the stable nuclei.

Figure 3: Segrè chart, in which the most studied nuclei are shown.

It can be observed that the most stable nuclei follow a trend, which consists in Z ≃ N
for A < 40. From A ≃ 40 on the ratio N/Z gradually increases until N/Z ≃ 1, 56 or
Z/A = 1/2, 5.

Page: 9
FNP

If the mass were represented in the third axis, the called “stability valley” would be
obtained:

From the study of this plot, the tables of nuclides and the isotopic abundance of the
stable nuclei the following properties can be inferred:

• The Pauli Exclusion Principle explains the numeric balance between protons and
neutrons, N/Z ≃ 1 for A < 40. However, for heavy nuclei (A > 40), the trend is
the relative increasing in the number of neutrons (N/Z ≃ 1, 56).

• There are much more stable nuclei of the even-even type. That means, there are
pairing forces. In particular, nuclei formed by several α particles ( 4 He, 8 Be, 12 C,
16
O) have a larger binding energy ε.
The four stable nuclei are: 21 H, 63 Li, 105 B and 147 N, all of them very light.
There is not a stable nucleus with 5 nucleons.

• The plot shows a large number of stable nuclei around the numbers Z or N = 2, 8,
20, 28, 50, 82 and 126. They are called “magic numbers” and sometimes 40 is also
included.
Nuclei with double magic numbers can be found as well, Z and N , for instance:
2 He, 8 O, 20 Ca and 82 Pb.
4 16 40 208

• Beyond the lead, Z = 82, the Coulombic repulsion breaks down the nuclei stability
and all the nuclei are unstable.

1.5.3 Nucleon separation energy


It is called “neutron separation energy” (Sn ), to the amount of energy that is needed
to remove a neutron from a nucleus. This process can be described as:
A
ZX + energy −→ A−1
ZX +n (23)

where it is assumed that some energy must be provided to the AZ X nucleus to remove the
neutron.
In terms of energy, the previous expression can be written in this way:

Ma (Z, N )c2 + Sn (Z, N ) = Ma (Z, N − 1)c2 + mn c2 (24)

If Sn (Z, N ) is isolated:

Sn (Z, N ) = −Ma (Z, N )c2 + Ma (Z, N − 1)c2 + mn c2 (25)

Page: 10
FNP

and comparing with (17) it follows that Sn (Z, N ) = −Qn (Z, N ).

The neutron separation energy Sn (Z, N ) can also be written as a function of the
binding energies. Starting from expression (25), if we add and subtract Ma (Z, N )c2 and
(N − 1)mn c2 in the right-hand side of the equation, together with equation (21), we can
arrive to:
Sn (Z, N ) = EL (Z, N ) − EL (Z, N − 1) (26)
The value of Sn (Z, N ) is always bigger for even N as a consequence of the parity force
and it ranges between 5 and 15MeV, dependent upon the nuclei.

The “proton separation energy” (Sp ) is defined in a similar way: Sp (Z, N ) =


−Qp (Z, N ). It is:
Sp (Z, N ) = −Ma (Z, N )c2 + Ma (Z − 1, N )c2 + Ma ( 1 H)c2 (27)
and as a function of the binding energies:
Sp (Z, N ) = EL (Z, N ) − EL (Z − 1, N ) (28)
Sp ranges between 1 and 14 MeV, depending on the nuclei.

There are peaks in the values of Sn and Sp for some values of N and Z, which cor-
respond exactly to the magic numbers. These energies follow the same trend than the
atomic ionization energy and reflects the existence of “shells”, as their values are high
when moving through some instances of Z and/or N , which coincides with the magic
numbers previously seen. The separation energies are also much larger for nuclei with
even N or Z and confirm the presence of the pairing force.
The pairing force can be determined separately for protons and neutrons and, in fact, this
is the obtained experimental value. For protons this energy is:
Pp (Z, N ) = Sp (Z, N ) − Sp (Z − 1, N ) (29)
while for neutrons:
Pn (Z, N ) = Sn (Z, N ) − Sn (Z, N − 1) (30)

1.5.4 Semiempirical mass formula


To calculate the atomic masses, Carl Friedrich von Weiszäcker proposed a method based in
the liquid-drop concept. In this model, it is assumed that the nucleus is an incompressible
drop, it is, of constant density and independent of A. Therefore, every nucleon has the
same binding energy, which is given by the following function, commonly known as “Bethe-
Weiszäcker formula” or “Semiempirical mass formula”:
EL (Z, N ) = αV A − αS A2/3 − αC Z(Z − 1)A−1/3 − αA (A − 2Z)2 A−1 + δ (31)
αV , αS , αC and αA are constants and δ is the pairing energy related to the stability of
the even-even nuclei compared to the odd-odd nuclei. Its value is:
 +αP A−1/2 even-even

δ= 0 even-odd (32)
−αP A −1/2
odd-odd

with αP a constant.

Page: 11
FNP

The Bethe-Weiszäcker formula is valid for nuclei with mass number A > 20 and cannot
explain the differences between binding energies of close nuclei. It is limited to reproduce
in a soft way the mass of the most stable nucleus among all the nuclei with the same mass
number A. It cannot predict masses far from the stability valley.

1.5.5 Mass parabola


Formulae (21) and (31) can be combined to express Ma (Z, N ) as a function of A and Z:
Ma (Z, N )c2 = Ama c2 − αV A + αS A2/3 + αA A − δ +
 

Ma ( 1 H)c2 − ma c2 − αC A−1/3 − 4αA Z+ (33)


 

αC A−1/3 + 4αA A−1


 

where the fact that A = Z + N has been employed so that it has been considered
Ma (Z, N )c2 ≡ Ma (Z, A)c2 .
If A = constant is done in equation (33), the functional form of Ma (Z, N ) (or Ma (Z, A))
is a parabola usually called “mass parabola”.

1.6 Nuclear Quantum Numbers


1.6.1 Wave function
Let’s now address the definition of the most employed quantum numbers in Nuclear
Physics. They are interesting because they characterise the properties of the nuclei and
sometimes it is possible to define them as a function of the quantum numbers of the
constituents nucleons.
From the quantum point of view, the state of a nucleus is given by a wave function which
depends on the spatial coordinates (⃗ri ), on the spin (⃗
si ) and the isospin (⃗ti ) of the A
nucleons. As these coordinates belong to independent spaces, the wave function can be
defined as the following product:
ψ(q1 , ..., qA ) = ψ(⃗
r1 , ..., r⃗A )χS (1, ..., A)ξT (1, ..., A) (34)
which must also be an anti-symmetric function under the exchange of two nucleons, as
the symmetrization principle postulates in the case of identical fermions.

If a model of individual particle in a central potential is assumed, it will be possible


to define, in the spatial part, states for each nucleon in an independent way:
ψ(⃗
r1 , ..., r⃗A ) = ϕ(⃗
r1 )...ϕ(r⃗A ) (35)
As for the H-atom, the functions will be separable:
ϕ(⃗
ri ) = Rni (ri )Yli mi (θi , φi ) (36)
in the radial Rn and the angular components, represented the latter by the spherical har-
monics Ylm .

The quantum numbers that define the state of each nucleon given by expression (36)
are n, l and m, it is, the principal quantum number, the orbital angular momentum and its
third component, respectively. Moreover, the total angular momentum (j) and its third
component (mj ) must be considered. They are obtained from ⃗j = ⃗l+⃗s and mj = −j, ..., j.

Page: 12
FNP

1.6.2 Nuclear spin and parity


Nuclei have well defined J P spin-parity. If it assumed that the nucleons that form the
nucleus move in a central potential, they will have an orbital momentum (⃗li ). Moreover,
the intrinsic angular momentum will be j⃗i = ⃗li + s⃗i . The nuclear total angular momentum
is the addition of all of them:
X A

J= j⃗i (37)
i=1

It is fulfilled, therefore, that if A is even, J⃗ is integer and if A is odd, J⃗ is half-integer. All


the nuclear states have a finite angular momentum and to specify it, the quantum vector
J⃗ is employed, its modulus is:
⃗ = ℏ J(J + 1) (38)
p
|J|
J is conserved in all the nuclear processes (“it is a good quantum number”). Due to
the Uncertainty Principle, J is unobservable, but this third component, according to a
quantization axis, can be measured and is equal to ℏmJ , where mJ can take any of the
(2J + 1) values in the range −J ⩽ mJ ⩽ J.
Finally, to mention that the even-even nuclei have J P = 0+ .

1.6.3 Nuclear isospin


The isospin is a new quantum number related to the nuclear force and explains the p − p
and n − n symmetry with respect to this interaction.
The proton and the neutron are considered the same object: the nucleon (N ). The nu-
cleon has a total isospin t = 1/2 and its projections are the proton with third component
t3 = +1/2 and the neutron with third component t3 = −1/2.
The N − N system can form a triplet state (T = 1) or a singlet state (T = 0) of isospin.
Each nucleus has an isospin T⃗ so that the third component is an operator which has
(2T + 1) eigenvalues (−T, ..., T ).

As a consequence of the rotational invariance of the Hamiltonian, in the isospin space


it will only depend on the total isospin T⃗ . For the nuclei it is fulfilled that:
|Z − N |
|T⃗ | ≥ (39)
2
with Tmax = A2 and T3 = 12 (Z − N ).
Considering the notation AZ XN , some examples of isospin multiplets are:
• Triplet (T = 1) of A = 10: 4 Be, 5 B5 , 6 C4 .
10 10 10

• Doublet (T = 1/2) of A = 7: 73 Li4 , 74 Be3 .


The electric charge Q and the third component of the isospin T3 are related to the mass
number A through the Gell-Mann-Nishijima formula (initially proposed for elementary
particles). In the case of nuclei with A nucleons, it can be written that:
A
Q = T3 + (40)
2
The electric charge is defined as the electromagnetic charge of a particle in |e| units
with e the electron charge. For instance: Q(proton) = +1, Q(neutron) = −1. Formula
(40) is also valid for nucleons, simply by substituting A with 1. DUDA: ¿Q(neutron) = 0?

Page: 13
FNP

1.7 Nuclear electromagnetic moments: dipole magnetic moment


The electromagnetic moments (charge, dipole moment, etc.) inform about the charge
distribution and nuclear magnetism originated in the movement of the nucleons. They
are important because they determine the behaviour of the nucleus in the presence of an
electromagnetic field.
The magnetic dipole moment of a nucleon is:

µ
⃗ = µ⃗l + µ⃗s (41)

where µ⃗l is the orbital dipole moment (due to the movement of the nucleon) and µ⃗s is the
spin dipole moment. They are defined in the following way:
µN ⃗ µN
µ⃗l = gl l and µ⃗s = gs ⃗s (42)
ℏ ℏ
gl and gs are the orbital and spin g factor respectively. µN is the nuclear magneton and
takes the value: µN = 2me p ℏ = 3.152 × 10−14 MeV/T.

In table (2) are collected the definition and value of the magnetic dipole moments of
the atomic constituents, namely, electrons, neutrons and protons.

electron µ⃗e = −gl µℏB ⃗l − gs µℏB ⃗s with gl = 1 and gs = 2.00232


neutron µ⃗n = gs µℏN ⃗s with gs = −3.826
proton µ⃗p = gl µℏN ⃗l + gs µℏN ⃗s with gl = 1 and gs = 5.586
Table 2: Definition and values of the magnetic dipole momentum µ
⃗ of the atomic con-
stituents.

Where µB is the Bohr magneton and its value is: µB = 2me e ℏ = 5.787 × 10−11 MeV/T.
Taking into account that ⃗s = ℏ2 ⃗σ , with ⃗σ being the Pauli matrices, the spin magnetic
dipole moments are:
µB µB ℏ µB µB
µe = −gs s = −gs = −gs = (−2.00232) ≃ −µB
ℏ ℏ 2 2 2
µN µN ℏ µN µN
µn = gs s = gs = gs = (−3.826) ≃ −1.913µN (43)
ℏ ℏ 2 2 2
µN µN ℏ µN µN
µp = gs s = gs = gs = (+5.586) ≃ +2.793µN
ℏ ℏ 2 2 2
Nuclei have a magnetic dipole moment:
µN ⃗
µ⃗J = gJ J (44)

with gJ the g factor of the nucleus. In a simple individual particle model, J⃗ can be
expressed as the addition of the angular moments j⃗i of the A nucleons.
The nuclei magnetic dipole moments are ranged between −2µN and 6µN .

Page: 14
FNP

2 Radioactivity and radiological protection


2.1 Introduction
Radioactivity is a phenomenon in which a nucleus emits one or several types of particles,
with the final nucleus being different from the initial (transmutation) or making a tran-
sition to a lower energy state of the same nucleus (deexitation).
Radioactivity can have a natural origin (radioactivity sources present in Earth crust or
cosmic rays originated from different astrophysical sources) or an artificial origin (particle
accelerator, etc.).

2.2 Types of disintegration


A nucleus can disintegrate in several ways and each one involves different fundamental
interactions. When a nucleus disintegrates, the energy of the resultant nucleus or nuclei
are lower (on the contrary case the disintegration would not occur).
The main types of disintegration that the nucleus can experience are:
• γ disintegration: AZ X∗ −→ AZ X + γ. This process consists in the transition between
two nuclear energy levels by the emission of a photon.
• β − disintegration: AZ X −→ Z+1 A
Y + e− + ν¯e . In this case, one electron of the nucleus
(not one from the atomic cloud) is emitted, originated from the conversion: n −→
p + e− + ν¯e (being ν¯e the electronic antineutrino).
• β + disintegration: AZ X −→ Z−1
A
Y + e+ + νe . It consists in the emission of a positron
coming from the conversion: p −→ n + e+ + νe (being νe the electronic neutrino).
• Electronic capture: AZ X + e− −→ Z−1 A
Y + νe . In this case, the nucleus captures a
“s” (from “s, p, d, f, etc.”) electron of the atomic cloud, giving rise to the reaction:
p −→ n + e+ + νe . (DUDA: ¿COINCIDE CON LA DE β + DISINTEGRATION?)
• α disintegration: AZ X −→ A−4
Z−2 Y + 42 He. It is the emission of a He−4 nucleus, also
called “alpha particle”.
• Neutrons emission: A
ZX −→ A−1
ZX + n. It consists in the emission of a neutron.
• Protons emission: A
ZX −→ A−1
Z−1 Y + p. In this case, a proton is emitted.
A′ A′′
• Spontaneous fission: AZ X −→ Z′ B+ Z′′ C. It consists in the fragmentation of the
nucleus in other ones lighter.

2.3 Radioactive disintegration law


Nuclei with access to lower energy states gradually disintegrate, that is, are unstable. The
radioactive disintegration obeys the following rule: the probability of disintegration in a
time t is independent of the value of t. Therefore, the probability of disintegration per
time unit in any time t is a constant, which will be called λ.
It follows that the probability of disintegration between t and t + dt is λdt. Obviously, in
increases with time, but it is independent of the time t in which we start to measure.

Page: 15
FNP

Finally, if we multiply that probability by the number of nuclei of the sample N


(DUDA: ¿QUÉ SIGNIFICA SAMPLE N ?, it is obtained that, between t and t + dt, the
charge in nuclei is on average dN = −λdt (the − represents that those nuclei disappear).
The solution of that differential equation is:

N (t) = N0 e−λt (45)

expression named Radioactive disintegration law, being N0 = N (t = 0).


If it is divided by the number of nuclei, it can be checked that for any single nucleus,
the probability of continuing to exist past a time t is proportional to e−λt . As a result,
the probability that a nucleus disintegrates right after the time t is the probability of
continuing to exist passed t(e−λt ) multiplied by the probability of disintegrating right
after the interval dt: λdt ⇒ p(t) = e−λt · λdt, this is a distribution normalised to 1.
[λ] = s−1 ⇒ [1/λ] ≡ mean lifetime, because it is obtained when the mean value of t is
carried out: R ∞ −λt
te dt 1
⟨t⟩ = R0∞ −λt = (46)
0
e dt λ
The mean value of t is usually represented by e:
1
⟨t⟩ = λ
=e

(these three magnitudes have dimensions of t).

On the other hand, after a time ln λ2 , the number of nuclei reduces to its half, so that
quantity is called: Half life, t1/2 = ln λ2 .

2.4 Graphical representation of the mass parabola


When the mass parabola is displayed (formula (33)), one obtains two configurations de-
pending on the number of nucleons:

• A odd ⇒ only one parabola can be found.

• A even ⇒ two parabolas, separated 28, with δ (DUDA: ¿SEPARATED 28, DELTA?)
the coefficient that gives the pairing energy of the mass parabola.

Examples of these two are:


A = 125: (
49 In, 50 Sn, 51 Sb, 52 Te, 53 I, 54 Xe, 55 Cs.

A = 128: 49 In, 51 Sb, 53 I, 55 Cs, 57 La.


50 Sn, 52 Te, 54 Xe, 56 Ba.

The nuclei with Z less than the most stable (that with the lowest mass) will disinte-
grate by β − disintegration and those with higher Z, by β + disintegration; processes that
will continue to happen until reaching the most stable nucleus of the mass parabola. This
set of stable nuclei against the β distribution are those that form the stability valley.

Page: 16
FNP

Figure 4: Mass parabola for A = 125 and A = 128.

2.5 Dosimetry and radiological protection


In the current legislation can be found the definition of several units which serve to
quantify the biological effects of radiation. Let us point out that, although starting from
high doses it is possible to predict the level of damage in tissues and organs, at low doses,
it is not possible to predict if the death of one or several cells will have effects on a tissue.
A dose of ionizing radiation which would be safe is not known, but the legislation imposes
certain thresholds of radiation that cannot significantly exceed the natural radiation dose
which people received (soil, food, building materials, cosmic rays, etc.).
Some units of radiation measurement are the following:

• The “activity” is the number of disintegration per second of a radioactive sam-


ple. In the International System of Units, it is measured in “becquerel”, 1 Bq =
1 disintegration/s, but the “curie” is usually employed: 1 Ci = 3, 7 × 1010 Bq.

• The “exposure” is the electric charge generated by an air mass unit. The “roentgen”
(R) is defined as the exposure needed to generate an electric charge of 1 esu in the
air, with 1 esu = 3, 33 × 10−10 C. The conversion to the International System of
Units is: 1 R = 2, 58 × 10−4 C/kg.

• The “absorbed dose” takes into account the deposited energy per mass unit, irre-
spective of the type of ionizing radiation. The unit which has been historically
employed is the “rad”: 1rad = 100 erg/g. In the International System of Units is
the “gray” (Gy) which corresponds to 1 J/kg, so that the conversion factor if: 1 Gy
= 100 rad.

• The “equivalent dose” considers that not all the radiations cause the same biological
damage and multiplies the absorbed dose by a “w” factor which takes into account
that biological effectiveness. The equivalent dose is measured in “roentgen equivalent
man” (rem), but the unit in the International System of Units is the “sievert” (Sv).
The conversion factor is: 1 Sv = 100 rem.

Page: 17
FNP

2.6 Protection measures


To reduce the absorbed dose three factors must be considered:

• The exposure time: the longer the time exposed to radiation, the larger the received
dose.

• The distance to the radioactive source.

• The existence of armour-plates.

Page: 18
FNP

3 Nuclear models and nuclear phenomenology


3.1 Introduction
The study of the nuclear structure is impossible to address in an exact way. At low energy
(for example, below 2 MeV), the nuclei can be considered as quantum systems consisting
of many nucleons that interact among them via a very complex potential. At high energy,
methods derived from Statistical Mechanics are employed.
When solving the Schrödinger Equation of the nucleus, two types of nuclear models are
considered:

• Microscopic or single particle models. In these models, the nuclear structure is


described as a function of the degrees of freedom of the nucleons. The wave function
of the system is of the form Ψ(q1 , ..., qA ), where qi represents the position r⃗i , the
spin s⃗i and the isospin ⃗ti of the nucleon i. In this way, the Hamiltonian is built
up adding the kinetic energies of the A nucleons and including terms based in two-
nucleon interactions through a two-body potential (three-body potential terms can
appear):
A
X ℏ2 2 1 X
Ĥ = − ∇i + V (i, j) (47)
i=1
2m 2 i,j

The form of the V (i, j) potential is not given by any theory, it is phenomenological
and, in general, it depends on the problem that has to be solved. In the cases A > 4,
it is difficult to solve and “effective” interactions are usually considered, in which
one nucleon interacts with an average nuclear potential. This concept gives rise to
the single particle models, easier to address and which has achieved many successes
in the shell model, for instance.

• Macroscopic or collective models. The degrees of freedom are now collective coordi-
⃗ = 1 PA r⃗i , center of mass of the nucleus, and Q
nates, such as R A i=1
⃗ = PA r2 Y20 (θi , ϕi ),
i=1 i
electric quadrupole moment, which, as can be seen, are expressed in terms of micro-
scopic coordinates and, therefore, they can be related with the macroscopic models.
The Hamiltonian is a function of the collective properties too.

3.2 Macroscopic collective terms


The collective models explain the global properties of the nuclei and its basic idea is to
describe the nucleus as a fluid in which it only concerns the collective movement of the
nucleons.

3.2.1 Liquid drop model


The liquid drop model assumes the nucleus to be an incompressible fluid. One of its more
remarkable prediction has been studied in equation (31): the semiempirical mass formula.

3.2.2 Fermi gas model


In the Fermi gas model, the ground states of the nuclei are described as a degenerated
Fermi gas. As in the case of the electrons, the nuclei obey the Fermi-Dirac statistics.
The binding force among nucleons is usually represented by a simple potential well with

Page: 19
FNP

different depth for protons and neutrons. As in addition, the protons have electric charge,
so their potential is different from that of the neutrons.

3.2.3 Even-even nuclei collective properties


The ground states of the even-even nuclei always have J P = 0+ , due to the pairing force
among the nucleons, whose pairing energy or binding energy of the pair is:
11, 2
P p ≃ Pn ≃ √ MeV (48)
A
and it ranges from about 3MeV for light nuclei to 0, 7MeV for A = 220.
From the experimental observations of this type of nuclei, it is deduced that studying two
kinds of collective movements is necessary:
• Vibrations around the spherical shape, which is that of equilibrium, for light nuclei,
A < 150.
• Rotations of non-spherical systems, called “deformed nuclei”, for heavy nuclei, 150 ≤
A ≤ 190 and A > 220.

3.2.4 Vibrational model


The Vibrational model describes collective movements of the nuclei, following the image
of the liquid drop. It is assumed that in equilibrium, the nucleus is a sphere of radius
R0 (it is known that many nuclei are deformed in their ground state). The nucleus is
a homogeneous fluid and its shape is described in terms of spherical harmonics, whose
coefficients are the amplitudes or shape parameters αλµ :
" ∞ X λ
#
X
R(θ, ϕ, t) = R0 1 + αλµ (t)Yλµ (θ, ϕ) (49)
λ=0 µ=−λ

3.2.5 Rotational model


It has been already checked that the nuclei with closed shells and the magic nuclei have
spherical symmetry. However, nuclei whose equilibrium shape is not spherical can be
found. This is a consequence of:
• The nuclear forces, which is a short-range and attractive reaction.
• The Coulomb force, which is a large-range and repulsive interaction.
• The centrifugal force of a spinning nucleus.
Let us assume a spinning nucleus with an angular velocity ω ⃗ and a total angular
momentum J. Its rotation kinetic energy is 2 Iω , where I is the momentum of inertia.
⃗ 1 2
2
As J = Iω, this energy can be also expressed as J2I . Taking the quantum mechanical
value of J,
⃗ the energy for a spinning object can be written as:
ℏ2
E= J(J + 1) (50)
2m
Increasing the quantum number J is equivalent to adding energy to the nucleus, so that
the excited states will be obtained. Such states form a sequence known as “rotational
band”.

Page: 20
FNP

3.3 Microscopic or single particle models


3.3.1 Introduction
The collective models are unable to predict the quantum numbers (spin-parity) of the
ground states of the nuclei with odd A. These nuclei present a series of characteristics
which are necessary to address with other kind of models. The spin of the nuclei with odd
A is half-integer. The parity can be + or −. Unlike the ground states of the even-even
nuclei, they have magnetic dipolar moment and, if its spin is bigger than 1/2, non-null
electric quadrupole moment.

3.3.2 Extreme single particle model


The “extreme” single particle model is a particular case of the single particle model and
is applied to nuclei with odd A. In the extreme single particle model it is assumed that
the nucleons are paired in a structure with J P = 0+ , in such a way that the properties of
the nucleus are given by those of the only non-paired nucleon. This nucleon will occupy
a state with well-defined orbital angular momentum ⃗l and spin ⃗s, and, therefore, its total
angular momentum will be ⃗j = ⃗l + ⃗s. The total angular momentum of the nucleus is
J⃗ = ⃗j and its spin, that of the nucleon. Moreover, the parity of the nuclear wave function
is P = (−1)l , with l the right value between two possible: l = j ± 1/2.

3.3.3 Shell model


Within the approach of the extreme single particle model, the shell model describes the
nuclear properties deriving from an effective interaction of a nucleon with a potential.
The nucleon-nucleon interactions are not considered, but the interaction of a constituent
with the average potential. Unlike the atoms, where the nucleus creates the potential in
which the electrons move, in the case of the nucleus there is not a protagonist which can
be identified with the nuclear field. All nucleons have a contribution and give rise to a
potential which must be inferred in an empirical way and that is, as a result, phenomeno-
logical.
It is already well experimentally established that the nuclei are very stable when the
number of nucleons coincides with the magic numbers:
Z = 2, 8, 20, 28, 50, 82, 126, ...
N = 2, 8, 20, 28, 50, 82, 126, ...
The properties of these nuclei are:
• They have high binding energy (EL ).
• They present abrupt variations in the energy of the first excited state 2+ of the
even-even nuclei.
• They have separation energies Sp and Sn bigger than the neighbour nuclei.
The “spherical” shell model is a particular case of the shell model and describes the
interactions of the nucleons with an average field of spherical symmetry, which represents
the nuclear average binding. The Hamiltonian is written as follows:
X −ℏ2 2
H= hc (ri ) , with hc (ri ) = ∇ + V (ri ) (51)
i
2µ i

Page: 21
FNP

being V (ri ) an effective central potential, Hartree-Fock like, which describes the mean
interaction of a nucleon with the rest of nucleons. It is a one-body term and it can be
checked that it is an approximation to the Hamiltonian considered in (47). This potential
does not include other interactions, as for instance, the pairing interaction.

Taking into account that the nucleon-nucleon interaction depends on the spin, to prop-
erly describe the problem, two more corrective terms must be added to the Hamiltonian:
hc (ri ) in (51), at least. Such terms are:
⃗s
• The spin-orbit interaction: a l·⃗
ℏ2
, where ⃗l and ⃗s are the orbital angular momentum
and the spin of the nucleon, respectively.
⃗2
• A correction proportional to ⃗l2 , which slightly shifts the energy values b ℏl 2 .
In this way, the Hamiltonian of the i-th nucleon becomes:
⃗l · ⃗s ⃗l2
h(ri ) = hc (ri ) + a + b (52)
ℏ2 ℏ2
If the potential considered in (52) is a Saxon-Wood potential, the magic numbers are
correctly reproduced and the theoretical results coincide with the experimental results.
The spin-orbit interaction increases the binding energy of a single nucleon when ⃗l and ⃗s
are parallel. The states of each nucleon split according to the value of the total angular
momentum: ⃗j = ⃗l + ⃗s, that is, all the states with the orbital angular momentum l > 0
give rise to two states because j = l ± 1/2.
Now the states are characterised by the three quantum numbers (n, l, j) and it is obtained
that:
j = l + 21 ,
   a
3 +2l
En,l,j = ℏω n + + bl(l + 1) (53)
2 − a2 (l + 1) j = l − 12
with b = −0, 03ℏω and a = −20A2/3 MeV.

In this way, employing the notation “nlj”, the order for filling the states are the
following:
1s1/2 , (1p3/2 , 1p1/2 ), (1d5/2 , 2s1/2 , 1d3/2 ), ...
and the levels distribution of figure (5) is obtained.
Given a nucleus with odd A = Z +N , in which Z or N is odd, but not both simultaneously,
the filling of the shells of figure (5) is performed as follows:
• The odd number is identified, Z or N (let us remind that if both are even, it follows
that J P = 0+ ).
• Nucleons are added to each level, from bottom to top, taking into account that each
level can be occupied by 2j + 1 nucleons, until all nucleons have been allocated.
• The quantum numbers of the last filled sublevel, named “valence subshell”, determine
the spin-parity J P of the entire nucleus: J = j, P = (−1)l .
It is worth clarifying that in the third step the number of nucleons has not been
specified because the only important aspect is that the number is odd. In the case of
3, 5 or more, the nucleons become paired and these pairs do not contribute to the J P
spin-parity.
In this conditions the magic numbers are reproduced:

Page: 22
FNP

z}|{
Z = 2, 8, 20, 28, 50, 82, 114 , 126, ...
N = 2, 8, 20, 28, 50, 82, 114, 126, |{z}
184 , ...

where two stability islands for the super-heavy nuclei with A = 114(Z) + 184(N ) = 298
and A = 126(Z) + 184(N ) = 310 are also predicted.

Figure 5: Energy levels structure.

Page: 23
FNP

4 The nucleon-nucleon interaction


4.1 Introduction
Deuterium, 21 H, discovered by Harold Clayton Urey in 1032, is one of the isotopes of
the hydrogen and its isotopic abundance is 1, 5 × 10−4 . The other isotope of hydrogen,
tritium, is unstable and its half-life is t1/2 = 12, 32 years. Hydrogen and its other two
isotopes play an important role in Nuclear Physics. Their nuclei are known by a different
name: proton, deuteron (d) and triton (t) for 11 H, 21 H1 and 31 H2 , respectively.
In this chapter, the study of the deuteron is addressed, p − n system, which is the most
simple state bound for the nuclear force. It is important because it informs about the
properties of that force. In Nuclear Physics, the constituents of the nuclei are the proton
and the neutron, which belong to the “hadrons” family. The term “hadron” comes from
the Greek “hadros”, which means “strong”, and it refers to the particles that feel the strong
interaction.

4.2 Deuteron properties and wave function


The deuteron is the only bound and stable state system and it consists of two nucleons,
(p, n). The rest of combinations do not form any bound state: (p, p), (n, n). It mass is
md = 1875, 7 MeV/c2 and its binding energy is EB = 2, 2 MeV. The spin-parity of the
deuteron is J P = 1+ and it has an isospin of T = 0.

As the deuteron is a system which consists of a proton and a neutron, with ⃗sp and
⃗sn being their spins respectively, moving with an angular momentum L,⃗ its quantum
numbers can be calculated as follows:
• Parity:
P (d) = P (p)P (n)P (L) (54)
• Total angular momentum:
J⃗ = ⃗sp + ⃗sn +L
|{z}
⃗ (55)
| {z }
1 0, 1

Deuteron only has one bound state and it is found from the experiments that J = 1.
As the spin contribution in (55) can only be 0 or 1, the orbital angular momentum L
must take the values 0, 1 or 2.
On the other hand, P (p) = P (n) = + by agreement (for nucleons J P = (1/2)+ . As P (L)
is the parity due to the orbital movement described by the spherical harmonics, it follows
that P (L) = (−1)L . Therefore, (54) can be rewritten as:
+1 = (+1)(+1)(−1)L (56)
where it follows that L can only have even values. As we have obtained that L can solely
have the values 0, 1 and 2, in order to fulfill both restrictions at the same time, it must
be: L = 0 and 2.
But the fact that L can only be 0 and 2 excludes the possibility that L = 1 in (55), so
that the only valid value for the total spin is J = 1. In other words, if J⃗ = ⃗sp + ⃗sn + L,

equation (55), it follows that:

Page: 24
FNP

J S L
1 0 1
1 1 0, 1 , 2
As a consequence of this reasoning, it is obtained that the deuteron is a mixture of the
two following states:

1st : L = 0, S wave ⇒ 3 S1 state


2nd : L = 2, D wave ⇒ 3 D1 state

where the notation followed in spectroscopy has been employed: 2S+1 LJ , in which the
letters S, P, D, ... represent the values L = 0, 1, 2, ...
The total spin J is conserved (“it is a good quantum number”), in this vase, J = 1.
However, the deuteron can be in any of the two possible states of L, L = 0 and L = 2. It
is found experimentally that the probabilities to find the deuteron in a S or D state are:
p(L = 0) = 96/100 and p(L = 2) = 4/100, in other words, it is more likely to find the
deuteron in the S state.

The wave function of the deuteron can be expressed in the following way:

|d⟩ = |ϕ(⃗r)⟩L |χ⟩S |ψ⟩T (57)

From the point of view of the isospin, proton and neutron are identical particles, as both
are also fermions, it can be gathered that |d⟩ is antisymmetric.
The spatial function |ϕ(⃗r)⟩L has the symmetry of the orbital angular momentum L. As
L is even (L = 0, L = 2), |ϕ(⃗r)⟩L must be symmetric and it can be written as:

|ϕ(⃗r)⟩L = aS | 3 S1 ⟩ + aD | 3 D1 ⟩ (58)

with a2S + a2D = 1.

In the Schrödinger equation, a central potential that is, a potential that only depends
on |⃗r|, would give rise only to the S component. As the D component is also present, the
potential cannot be uniquely central.
In the case of the spin function |χ⟩S , the fact that S = 1 implies that ⃗sp and ⃗sn are
parallel, and, therefore, it is a symmetric function.

As a consequence of all of these facts, the isospin wave function |ψ⟩t must be antisym-
metric, representing in principle a singlet or a triplet. But it cannot be a triplet, because
in that case, one would fin (p, p) and (n, n) bound states, and it is not so. Therefore, |ψ⟩T
can only be an antisymmetric singlet.
    
1 1 1 1 −1 1 −1 1 1
|T = 0, T3 = 0⟩ = √ , , − , , (59)
2 2 2 1 2 2 2 2 2 1 2 2 2

where the nucleons have been denoted as |t, t3 ⟩ which represents a well defined isospin
state.

Page: 25
FNP

4.3 Yukawa potential


Hideki Yukawa employed the quantum idea that the nucleon-nucleon interaction is due to
the exchange of a particle between them (1934, Physics Nobel Prize in 1949). The diagram
that displays this idea can be seen in figure (6), usually called “Feynmann diagram”.

Figure 6: Exchange of a Yukawa meson diagram.

In this way, Yukawa introduces the idea that the nuclear interactions are mediated by
the exchange of a particle called “meson”. The root of this word comes from the Greek
“meso”, which means “middle”, and was named as such because all the masses of this kind
of particles took intermediate values between the lightest and the heaviest (DUDA: EN
CLASE PUSO the lighter and the heavier (?). All the mesons are bosons (integer spin).
This set of ideas gave rise to the called OPEP model, which is the acronym of “One pion
exchange potential”, with the pions being a type of mesons.
The Yukawa potential describes properly the problem at distances larger than approxi-
mately 1 fm and it is expressed as:

gs e−r/r0
V (r) = (60)
4π r
where gs is the coupling constant of the nuclear force and r0 = mc ℏ
, with “m” the mass of
the emitted boson.
gs2
The dimensionless coupling constant of the strong force is αs = 4πℏc ≃ 1 and when it is
compared with the dimensionless coupling constant of the electromagnetic force, which
2
is the fine structure constant, α = 4πϵe 0 ℏc ≃ 137
1
, it is understood the name of “strong” force.

4.4 Nucleon-nucleon potential


From that studied up to now about the nucleon-nucleon potential, the following conclu-
sions can be drawn.

• At short distances there is an impenetrable core and, therefore, repulsive. This fact
will give rise to a zone of positive potential with a strong upwards increase as r
approaches to 0.

• The Yukawa potential provides a reasonable description for larger distances that
approximately 1 fm.

Page: 26
FNP

From all these considerations, one can conclude that an attractive intermediate zone
is present, deeper than that of the Yukawa potential, where the bound states are found.
This potential is displayed in figure (7), for the n − n case.

Figure 7: Nuclear potential.

The difference of this potential with that of n − p and p − p is that the latter present a
Coulomb barrier. In the intermediate zone, the energy of any bound state must fulfill:
−VO < E < 0 (V0 > 0, if E > 0 the process would be a collision).

Page: 27

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy