The Discovery of The Atomic Nucleus: Doi:10.1088/978-0-7503-1140-3ch1

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 16

Chapter 1

The discovery of the atomic nucleus

Nuclear physics began on 1 March 1896 with the discovery of radioactivity by Henri
Becquerel. Becquerel discovered that uranium salts (such as K 2UO2 (SO4)2·2H2O)
generated black tracks on photographic plates. The radiation responsible was first
interpreted as penetrating UV-radiation since Wilhelm Röntgen had discovered
x-rays the year before. Becquerel called his radiation ‘rayons uraniques’. The word
‘radioactivity’ was introduced later by Marie Curie. Ernest Rutherford and
others discovered that radioactivity consisted of at least three components: (i) α-
radiation, which is easily absorbed and deflected by magnetic fields, (ii) penetrating
β-radiation (which Becquerel had initially observed), which is deflected in the
opposite direction, and (iii) γ-radiation, which is not deflected by magnetic fields. α-
and β-rays therefore had opposite electric charges, while γ-radiation was
electrically neutral.
Rutherford described the energy released as ‘atomic’ energy1. Since atoms were
perceived as elementary building blocks 2 the origin of the release of energy was not
clear: was the theorem of energy conservation violated or could the atom after all be
decomposed into smaller fragments?
More radioactive atoms were soon discovered, including polonium and
radium (Pierre and Marie Curie, 1898, see figure 1.1). In 1909 Rutherford and
Geiger observed the characteristic spectral lines of helium in a glow-lamp irradiated
by a radium source (helium had been discovered in 1895 by Norman Ramsey in
uranium ore). They concluded that helium atoms were α particles which had lost
their positive charges!

1
This word was used until the 1950s, after the development of the ‘atomic’ bomb. The more appropriate word
‘nuclear’ would have suggested to the media a relationship with the cell nucleus, i.e. with biology (after
L Groves, head of the Manhattan Project).
2
The word ‘atomos’ was introduced by Democritus. For the Greeks the fundamental building blocks (atoms)
were the four elements earth, air, fire and water. The belief that water was elementary lasted until Lavoisier
(1743–94). The atom as the fundamental unit of chemical substances was introduced by Dalton (1808). He
discovered that the mass of a chemical substance was given by the sum of the masses of its components.
Prout (1815) also noticed that the ratio of the mass of an atom to that of the most elementary one —hydrogen—
was approximately an integer.

doi:10.1088/978-0-7503-1140-3ch1 1-1 ª IOP Publishing Ltd 2015


Nuclear and Particle Physics

Figure 1.1. Pierre and Marie Curie in their laboratory. Image from Wikimedia Commons.

J J Thomson was convinced, following his discovery of the electron in 1897, that
the atom consisted of a ‘pudding’ of electrons and positive charges, where (strangely
enough) only electrons contributed to the mass of the atom. Hydrogen thus
contained approximately 2000 electrons. This view was revised in 1906 once the
rate for the scattering of x-rays by shell electrons had been been measured and
compared with the so-called Thomson cross section. Hydrogen was found to have
1–2 electrons.
After the discovery of the atomic nucleus in 1911 by Rutherford, Geiger and
Marsden, which we shall describe in detail, the atom was modelled as a hard core
with A positive charges and (A − Z) electrons (total positive charge Ze) surrounded
by a cloud of electrons with charge −Ze. The electrons in the core were considered to
be responsible for the β-radiation. (The neutron was discovered in 1932 and the
theory of β-decay as the transformation of a neutron into a proton (n → p) was
formulated in 1934 by Enrico Fermi.) The simplest atom, hydrogen, consisted of one
unit of positive charge, which Rutherford called the ‘proton’, and one shell electron.
The planetary model of the atom as we know it today was introduced by Niels Bohr
in 1913. The first artificially produced nuclear reaction (α14N→p17O) was
observed in 1914 by Rutherford and Marsden. A comprehensive account of the early
days of nuclear and particle physics can be found in [1].

1-2
1.1 Rutherford scattering
Hans Geiger and Ernest Marsden, two employees of Rutherford, were investigating
α scattering off thin films at the University of Manchester. The apparatus (figure 1.2)
consisted of an evacuated chamber, a collimated radium source, R, and a ZnS
scintillator, S, which could be swung around a thin gold or silver foil, F, mounted in
the centre of the chamber. The kinetic energy of the α particle was T = 4.78 MeV.
The scattered α particles produced flashes of light in the ZnS scintillator which
could be observed using the microscope, M. The number of flashes for a given
measurement time is shown in table 1.1 as a function of scattering angle θ. At small
angles the collimation of the source was increased since the human eye cannot
resolve counting rates exceeding ∼90/min. Geiger and Marsden determined that the
rate decreased with increasing scattering angle, following the angular distribution
1/sin4 θ 2. Surprising was the observation of backward scattered α particles (θ > 90°).
This was only possible if the atom was made up of a hard heavy core in which the
electric charge was concentrated, the nucleus3.
Let us now describe the scattering process quantitatively (figure 1.3) by making
the following (modern) assumptions:
1. The nucleus is much heavier than the projectile (mass m, velocity ) so that
the recoil energy of the nucleus is negligible (that is the velocity v of the
projectile after the collision is v′ = v). This does not apply to free atoms but
to atoms that are bound in a crystal lattice.
2. The core and the projectile are structureless (point-like) with charges +Ze
and ze, respectively, and have spin 0.
3. The interaction is fully described by the electromagnetic interaction
(Coulomb force).
4. The nucleus is not excited in the scattering process (elastic scattering) and the
projectile does not lose energy by interacting with shell electrons since m is
much larger than the mass of the electron.

Figure 1.2. The apparatus used by Rutherford, Geiger and Marsden (after [2]).

3
According to Rutherford the phenomenon was like ‘shooting a cannon ball at a piece of tissue paper and have
the ball bouncing back at you’.
Table 1.1. Number of counts N as a function of scattering angle θ.

Silver Gold
1
θ [°]
θ2
N N sin
2
4 θ
N N sin
2
4 θ

sin4
150 1.15 22.2 19.3 33.1 28.8
135 1.38 27.4 19.8 43.0 31.2
120 1.79 33.0 18.4 51.9 29.0
105 2.53 47.3 18.7 69.5 27.5
75 7.25 136 18.8 211 29.1
60 16.0 320 20.0 477 29.8
45 46.6 989 21.2 1435 30.8
37.5 93.7 1760 18.8 3300 35.3
30 223 5260 23.6 7800 35.0
22.5 690 20300 29.4 27300 39.6
15 3445 105400 30.6 132000 38.4

30 223 5.3 0.024 3.1 0.01


4
22.5 690 16.6 0.024 8.4 0.01
2
15 3445 93.0 0.027 48.2 0.01
4
10 17330 508 0.029 200 0.0115

Figure 1.3. Rutherford scattering.

Let v⊥ = ωr be the component of the velocity perpendicular to r (figure 1.3).


One obtains for each point on the trajectory from the conservation of angular
momentum L⃗ that
dϕ dϕ vb
L = r ⃗× = mvb = mv⊥r = mωr2 = m r 2 ⇒ = . (1.1)
mv
dt dt r2
The impact parameter b is the distance between the incident projectile at asymptotic
distances and the beam axis. The scattering angle is denoted by θ, and Kn is the
force projected in the direction n⃗ . The projection of the asymptotic momenta on
n⃗ is
−mv cos(π − θ)/2 = −mv sin θ /2 before, and mv + sin θ /2 after the collision. Thus
along n⃗ the momentum changes by

Δp = 2mv sin
θ
+∞
zZe2
2
= ∫ ∫
dt = −∞
4πϵ0r2
cos ϕdt. (1.2)
Kn
From (1.1) we have that
r
dt = 2 dϕ , (1.3)
vb
and hence
zZe zZe θ
Δp = ∫ 2
(π −θ)
2 cos ϕdϕ = 22 cos . (1.4)
−(π −θ)
4πϵ0vb 4πϵ0vb 2
2

Using (1.2) we obtain


θ ⎤
tan 1

= zZe 2 ⎥ . (1.5)

2 ⎣ 4π ϵ0mv2 ⎦ b
–_–,
K
One finds a one-to-one relation between the impact parameter b and the scattering
angle θ. For large (respectively small) distances from the axis,

b → ∞ ⇔ θ → 0,
b → 0 ⇔ θ → π. (1.6)

Consider now a sphere with unit radius (figure 1.4): incoming particles which
fly through the ring with area

dσ = 2πb∣db∣ F
i
g
u
r
e
1
.
4
.
T
h
e
u
nit sphere with unit radius.
(1.7)
are scattered between θ and θ + dθ into the solid angle
dΩ = 2π sin θdθ = 2πd(−cos θ). (1.8)

The area dσ is referred to as the differential cross section. From (1.5) one obtains
⎛ ⎞ ⎛ ⎞

∣db∣ = d⎜⎜ K ⎟ = ⎜⎜
K ⎟ 1
2 θ ⎟
θ 2
dθ, (1.9)
⎝ tan2 ⎝ tan 2 2 cos2

hence with (1.7)


⎛ 2
⎜ K ⎞⎟ 1
dσ = 2π ⎝ ⎠ dθ. (1.10)
⎜ tan3 θ ⎟ 2 cos2 θ
2 2

Dividing dσ by dΩ from (1.8) one obtains


1 ⎛ zZe2 ⎞
2
dσ ⎛ 2
1 K 1
= ⎜⎜ 2 ⎟⎞
dΩ K 3θ ⎟
θ 2 = = ⎜ ⎟ 4 , (1.11)
4 θ θ
⎝ tan 2cos sin 4 sin 16 ⎝ 4πϵ0T sin
⎠2 θ
2 2 ⎠ 2

where T = mv2 /2 is the kinetic energy of the projectile. Let us introduce the
dimensionless fine-structure constant

e2 =1
α≡ , (1.12)
4πϵ0 ℏc137.036

as well as the natural units

ℏ = c = 1. (1.13)
Equation (1.11) then takes the simple form known as the Rutherford formula

dσ = z2Z2α21 , (1.14)
dΩ16T 2 sin4 θ
2

which is the differential cross section for scattering into the solid angle dΩ. As we
shall see, formula (1.14) is a measure of the probability for the projectile to be
scattered between θ and θ + dθ and therefore describes the angular distribution of
the scattered particles.

1.1.1 Remarks on units


Energies in nuclear and particle physics are always specified in eV (keV, MeV, GeV,
TeV), where 1 eV is the energy increment of an elementary charge in a potential of
1 V: 1 eV = 1.6 × 10−19 J. How do we then get from (1.14) for dσ the dimension of a
surface (e.g. cm2) when T is given in MeV? With our choice of units (1.13) cm−1
is redefined in MeV:
ℏc = 197.3 MeV fm = 1, (1.15)
where 1 fm = 10−13 cm. Hence

1 MeV ≡ 1fm−1
. (1.16)
197.3

In a similar manner with c = 1 one obtains


1 s = 2.998 × 1023 fm (1.17)
and

1 MeV = 1.519 × 1021 s−1. (1.18)

Hence in all calculations and c can be set equal to 1. The units obtained in the final
result can then be converted into cm, s or MeV using (1.16) or (1.17). Energies are
usually given in MeV and masses in MeV c−2 since E 2 = c2p2 + m2c4 (or in MeV
with c = 1). Momenta are normally expressed in MeV c−1 to avoid confusion
between energy and momentum.
Furthermore, in (1.12) the electric charge e2 has been replaced by the dimen-
sionless α where
e2 = 4πϵ0α, (1.19)
since we used SI units for the Coulomb force. Sometimes the Coulomb force is
expressed as e2/4πr2 (in the so-called Heaviside–Lorentz units with ϵ0 = 1). The
conversion then reads
e2 = 4πα
(1.20)
for the result to agree with (1.14). Alternatively, in ESU the Coulomb force is
expressed as e2/r2. The correct substitution is then
e2 = α
(1.21)
to obtain (1.14). It is therefore mandatory to determine in which units the Coulomb
force was written before replacing e2 by α. It must be emphasized that α always
assumes the numeric value 1 in any system of units.
≃ 137
1.1.2 Interpretation of the differential cross section
Consider a beam of cross section F impinging on a target (thickness ℓ) with I0
incident particles per second (figure 1.5). Let us assume that an incident particle is
scattered by (or interacts with) a target nucleus whenever it crosses an effective
surface σ. This surface (total cross section) depends on the incident energy T,
the types of projectile and target, and the nature of the interaction.
Figure 1.5. Definition of the cross section σ.

The scattering probability for one incident particle and one target nucleus is
σ /F . On the other hand, the number of scattering centres crossed by the beam
is nℓF , where n is the number of nuclei per unit volume. Assuming that ℓ is small
enough so that the surfaces σ do not overlap one obtains the number of scatterings
per second:
σ
I= nℓFI0 = I0σnℓ, (1.22)
F

with
ρL
n , (1.23)
= A

where ρ is the density, L is Avogadroʼs number and A is the atomic weight.


Cross sections σ are usually expressed in units of ‘barn’:

1b ≡ 10−24 cm2 . (1.24)

(Cross sections for slow neutrons on uranium are very large, they are ‘as big as a
barn’.) The number of scattered particles per second into the solid angle dΩ is then
dI dσ dI dσ
dΩ = I0nℓ dΩ ⇒ = I0nℓ . (1.25)
dΩ dΩ dΩ dΩ

The differential cross section dσ (dimension b sr−1) describes the angular distribution

of the scattered particles and depends on the physical process. The scattered rate R
into a detector of surface S located at a large enough distance d from the target is
then with ΔΩ ≃ S /d 2:
⎛ dσ ⎞ S
R = I0 nℓ ⎜ ⎟ . (1.26)
⎝ dΩ ⎠d
2
1.1.3 Remarks on Rutherford scattering
The differential cross section (1.14) for Rutherford scattering decreases as 1/sin4 2θ
with increasing angle θ, in accordance with the experimental observations of Geiger
and Marsden (figure 1.2). The atom thus has a hard core with charge +Ze. A series
of remarks is in order here:
1. The differential cross section increases with Z2 for fixed angle θ.
2. dσ
dΩ become infinite when θ → 0. That should not bother us since for very
small scattering angles the impact parameter b exceeds the radius of the outer
electronic shell. The charge of the core is then shielded and the factor Ze → 0.
3. Nevertheless, when scattering off a bare (completely ionized) nucleus,
the differential cross section diverges for θ → 0. Furthermore

σ= ∫ dσ
2π sin θdθ → ∞, (1.27)

so that the total cross section σ for Rutherford scattering is actually not
defined. This is a property of the electromagnetic interaction which is related
to its infinite range. In a sense the choice of electromagnetic scattering to
define the cross section σ in figure 1.5 was inappropriate. However, for
interaction potentials decaying faster than 1/r (such as nuclear potentials) σ
remains finite.
4. The distance of closest approach to the centre of the nucleus is (see figure 1.3)
zZα ⎛⎜ 1 ⎟⎞
D= 1+ , (1.28)

2T ⎜ ⎟
⎝ sin θ
2

the derivation of which is left as an exercise for the reader. The minimum is
reached for θ = 180° (backwards scattering):
zZα
Dmin = . (1.29)
T
The distance of closest approach decreases with decreasing Z and
increasing kinetic energy T. When the radius of the nucleus becomes larger
than D one expects a deviation from the Rutherford formula (1.14) because
the assumption of a point-like nucleus collapses. A good approximation
for the nuclear radius is a ∼1.2 fm A13 (see chapter 2). For instance, for
gold (Z = 79, A = 197) a ∼7 fm and for T = 4.78 MeV, z = 2, the minimum
distance is Dmin = 48 fm (to convert the units use (1.16)). Geiger
and Marsden were therefore lucky: with much higher kinetic energies
and/or lighter targets their data would not have agreed with (1.14). From
the Rutherford formula one can derive an upper limit of 48 fm for the
radius of the gold nucleus.
In 1921 Chadwick and Bieler studied α-proton scattering with a hydrogen
target (Z = 1). The measured differential cross section departed significantly
from the Rutherford formula. (In this case the recoil energy of the proton
would
have to be taken into account in the derivation of the scattering
formula.) Models including finite charge distributions in the α particle and in
the proton could not describe the data. The reason for the discrepancy was
the onset of the strong (nuclear) interaction which only acts at very small
distances (∼1 fm). A satisfactory explanation for this new force was provided
in 1935 by Yukawa
who introduced a hypothetical particle, the pion (π) (section 7.3.2). However,
the year 1921 can be considered as the actual birth year of particle physics.
The radii of many nuclei (and of the proton) were measured during the
1960s. To avoid complications arising from the finite size of the α particle,
point-like projectiles were used, namely electrons. This requires a relativistic
derivation of the differential cross section—including the influence of the
electron spin—for which the Dirac equation is required (chapter 15).
Intimations of this will be given in the next chapter.
5. Could the large angle scattering observed by Geiger and Marsden be due
instead to multiple scattering in the target? Consider the problem of the one-
dimensional random walk with m forward or backward steps of lengths
Di = ±1, each with 50% probability. After one step the average path length is
〈D1〉 = 0, but the mean square distance is

D12 = 1. (1.30)

After i + 1 steps the mean square distance is with Di+1 = Di ± 1


Di2 1 = Di2 + 1 ± 2Di ⇒ Di2 1 = Di2
+ 1. (1.31)
+ +

Hence after m steps the mean square distance is


2
Dm = m ⇒ D m2 m. (1.32)
=

We can model multiple scattering as a series of increments or decrements θi


in a two-dimensional plane. From the analogy with the random walk we then
expect the mean scattering angle

θi2 = m ∝ℓ . (1.33)

Rutherford and co-workers found, however, that the scattering rate at large
angles increased proportionally with target thickness ℓ rather than with ℓ ,
in accordance with formula (1.25).
6. It is easy to show that the Rutherford formula (1.14) also applies to a
negative charge ze of the projectile (attractive Coulomb force, θ → −θ in
(1.5)). In this case the distance of closest approach becomes
zZα ⎜⎛ 1 ⎞⎟
D= −1 + →0 for θ → 180°. (1.34)

2
2T ⎜ ⎟
⎝ sin θ

References
[1] Pais A 1986 Inward Bound (New York: Oxford)
[2] Geiger H and Marsden E 1913 Phil. Mag. 25 604

You might also like