Download as pptx, pdf, or txt
Download as pptx, pdf, or txt
You are on page 1of 195

Nuclear pHYSICS

Gintaras ValiulisProf. Gintaras Valiulis


(Vilnius University, Laser Research Center)
INTRODUCTORY TERMINOLOGY
• We date the origin of nuclear physics from Becquerel's discovery of radioactivity in 1896 or Rutherford's hypothesis of
the existence of the nucleus in 1911, it is clear that experimental and theoretical studies in nuclear physics have played
a prominent role in the development of twentieth century physics.
• The nuclei are characterized by the total amount of positive charge in the nucleus and by its total number of mass units.
The net nuclear charge is equal to + Ze, where Z is the atomic number and e is the electron charge.
• The fundamental positively charged particle in the nucleus is the proton, which is the nucleus of the simplest atom,
hydrogen. A nucleus of atomic number Z therefore contains Z protons, and an electrically neutral atom therefore must
contain Z negatively charged electrons. Since the mass of the electrons is negligible compared with the proton mass (mp
≈ 2000me), the electron can often be ignored in discussions of the mass of an atom.
• The mass number of a nuclei, indicated by the symbol A, is the integer nearest to the ratio between the nuclear mass and
the fundamental mass unit, defined so that the proton has a mass of nearly one unit. For nearly all nuclei, A is greater
than Z, in most cases by a factor of two or more. Thus there must be other massive components in the nucleus.
• Before 1932, it was believed that the nucleus contained A protons, in order to provide the proper mass, along with A - Z
nuclear electrons to give a net positive charge of Ze. However, the presence of electrons within the nucleus is
unsatisfactory for several reasons:
INTRODUCTORY TERMINOLOGY

In addition to the different penetrating powers of α-, β-, and γ-radiation, other properties were used to identify these mysterious
radiations, such as the differing deflections that the three radiations undergo in electric or magnetic fields. Alpha radiation was
known to possess a positive charge, because it would be deflected toward the negative electrode in an electric field potential,
while beta particles were known to be negatively charged due to their deflection in the opposite direction toward the anode or
positive electrode. Gamma radiation would not undergo any deflection, as illustrated in Figures above. 
INTRODUCTORY TERMINOLOGY
• The nuclear electrons would need to be bound to the protons by a very strong force, stronger even than the Coulomb
force. Yet no evidence for this strong force exists between protons and atomic electrons.
•  If we were to confine electrons in a region of space as small as a nucleus ( ∆x ~ 10-14 m), the uncertainty principle
would require that these electrons have a momentum distribution with a range ∆p ~ ħ/∆x = 20 MeV /c. Electrons that
are emitted from the nucleus in radioactive β decay have energies generally less than 1 MeV; never do we see decay
electrons with 20 MeV energies. Thus the existence of 20 MeV electrons in the nucleus is not confirmed by
observation.
• The total intrinsic angular momentum (spin) of nuclei for which A - Z is odd would disagree with observed values if A
protons and A - Z electrons were present in the nucleus. Consider the nucleus of deuterium ( A = 2, Z = 1), which
according to the proton-electron hypothesis would contain 2 protons and 1 electron. The proton and electron each have
intrinsic angular momentum (spin) of ½, and the quantum mechanical rules for adding spins of particles would require
that these three spins of ½ combine to a total of either or . Yet the observed spin of the deuterium nucleus is 1.
• Nuclei containing unpaired electrons would be expected to have magnetic dipole moments far greater than those
observed. If a single electron were present in a deuterium nucleus, for example, we would expect the nucleus to have a
magnetic dipole moment about the same size as that of an electron, but the observed magnetic moment of the deuterium
nucleus is about 1/2000 of the electron's magnetic moment.
INTRODUCTORY TERMINOLOGY
•  Of course it is possible to invent all sorts of ad hoc reasons for the above arguments to be wrong, but the necessity for
doing so was eliminated in 1932 when the neutron was discovered by Chadwick. The neutron is electrically neutral and
has a mass about equal to the proton mass (actually about 0.1% larger). Thus a nucleus with Z protons and A - Z
neutrons has the proper total mass and charge, without the need to introduce nuclear electrons. When we wish to
indicate a specific nuclear species, or nuclide, we generally use the form , where X is the chemical symbol and N is the
neutron number, N=A - Z.
• Neutrons and protons are the two members of the family of nucleons. When we wish simply to discuss nuclear particles
without reference to whether they are protons or neutrons, we use the term nucleons. Thus a nucleus of mass number A
contains A nucleons.
•  Nuclides with the same proton number but different neutron numbers are called isotopes; for example, the element
chlorine has two isotopes that are stable against radioactive decay, 35Cl and 37Cl. It also has many other unstable
isotopes that are artificially produced in nuclear reactions; these are the radioactive isotopes (or radioisotopes) of Cl.
• It is often convenient to refer to a sequence of nuclides with the same N but different Z; these are called isotones. The
stable isotones with N = l are 2H and 3He. Nuclides with the same mass number A are known as isobars; thus stable 3He
and radioactive 3H are isobars.
UNITS AND DIMENSIONS
• In nuclear physics we encounter lengths of the order of 10 -15 m, which is one femtometer (fm). This unit is colloquially
known as one fermi, in honor of the pioneer Italian-American nuclear physicist, Enrico Fermi. Nuclear sizes range from
about 1 fm for a single nucleon to about 7 fm for the heaviest nuclei.
• The time scale of nuclear phenomena has an enormous range. Some nuclei, such as 5He or 8Be, break apart in times of
the order of 10-20 s. Many nuclear reactions take place on this time scale, which is roughly the length of time that the
reacting nuclei are within range of each other's nuclear force. Electromagnetic ( γ) decays of nuclei occur generally
within lifetimes of the order of 10-9 s (nanosecond, ns) to 10-12 s (picosecond, ps), but many decays occur with much
shorter or longer lifetimes. α and β decays occur with even longer lifetimes, often minutes or hours, but sometimes
thousands or even millions of years.
• Nuclear energies are conveniently measured in millions of electron-volts (MeV), where 1 eV = 1.602 X 10-19 J is the
energy gained by a single unit of electronic charge when accelerated through a potential difference of one volt. Typical β
and γ decay energies are in the range of 1 MeV, and low-energy nuclear reactions take place with kinetic energies of
order 10 MeV. Such energies are far smaller than the nuclear rest energies, and so we are justified in using nonrelativistic
formulas for energy and momentum of the nucleons, but β-decay electrons must be treated relativistically.
• Nuclear masses are measured in terms of the unified atomic mass unit, u (or a.m.u), defined such that the mass of an
atom of 12C is exactly 12 u. Thus the nucleons have masses of approximately 1 u. In analyzing nuclear decays and
reactions, we generally work with mass energies rather than with the masses themselves. The conversion factor is 1 u =
931.502 MeV, so the nucleons have mass energies of approximately 1000 MeV. The conversion of mass to energy is of
course done using the fundamental result from special relativity, E = mc 2; thus we are free to work either with masses or
energies at our convenience, and in these units c2 = 931.502 MeV/u.
• 1 AMU = (1/12) m(12C) = 1.6605 ×10−27 kg .
NUCLEAR BINDING ENERGY
The binding energy of a nucleus is the difference in mass energy between a nucleus and its constituent Z protons and N
neutrons:
𝐸 𝐵 ( 𝑍 , 𝐴 )=[ 𝑍 𝑚𝑝 + 𝑁 𝑚𝑛 − 𝑀 ( 𝑍 , 𝐴)  ] 𝑐 2
where: mp is the mass of proton, mn is the mass of neutron, M(Z,A) is the mass of nucleus. With the masses generally
given in atomic mass units, it is convenient to include the unit conversion factor in c 2, thus: c2 = 931.50 MeV/u.
Thus for masses given in energy units we can simply write:

Usually we be dealing with atomic masses. Grouping the Z proton and electron masses into Z neutral hydrogen atoms ,
as + we can rewrite equation for binding energy as
𝐸 𝐵 ( 𝑍 , 𝐴 )=𝑍 𝑀 𝐴 (1,1)+ 𝑁 𝑚𝑛 − 𝑀 𝐴 ( 𝑍 , 𝐴)

where: MA(1,1) is the mass of Hydrogen atom, mn is the mass of neutron, MA(Z,A) is the mass of atom.

For practical use we introduce the mass defect for nucleus .

Since we have , for the binding energy we obtain :


or
NUCLEAR BINDING ENERGY
The masses and mass defects for proton, neutron and α-particle (in MeV and atomic mass units) are:

mp=938.2796 MeV = 1.0073 a.m.u, ∆p=6.7761 MeV = 0.0073 a.m.u;


mn=939.5731 MeV = 1.0087 a.m.u, ∆p=8.0715 MeV = 0.0087 a.m.u;
mα=3727.4053MeV = 4.0015 a.m.u, ∆α 1.4028 MeV = 0.0015 a.m.u

The binding energy of α-particle:

Usually in the atomic mass tables one can find not the mass defect of nucleus ∆(Z,A), but rather the mass defect of atom

∆A(Z,A) = MA(Z,A)-A= ∆(Z,A)+Zme .

Example: for the mass defect of the we should add the masses of 2 electrons:
∆A(2,4) = ∆(2,4) + 2me = 1.4025 Mev +2×0.5110 Mev=2.4248Mev

The equation we could rewrite as:

RULE FOR BINDING ENERGY:


Masses (or Mass defects) of the FINAL state – Masses (or Mass defects) of the INITIAL state
NUCLEAR BINDING ENERGY
EXAMPLES:

1) Binding energy for proton (proton extraction energy):

Since

Thus, finally:

2) Binding energy for neutron (neutron extraction energy):

3) α-particle extraction energy (α-particle binding energy in nucleus):

RULE:
BINDING ENERGY of some fraction of nucleus = binding energy of INITIAL particle – binding energies of the
constituents of the FINAL state, i.e.the rule is opposite as for masses
NUCLEAR BINDING ENERGY
Some Mass Defects and Separation Energies
𝑬 𝒏𝑩 𝑬 𝒑𝑩

Fission

Fusion

Binding energy per nucleon

Since the binding energy increases more or less linearly with A, it is general practice to show the average binding energy per
nucleon, , as a function of A. Figure shows the variation of with nucleon number. Several remarkable features are immediately
apparent.
First of all, the curve is relatively constant except for the very light nuclei. The average binding energy of most nuclei is, to
within 10%, about 8 MeV per nucleon.
Second, we note that the curve reaches a peak near A = 60, where the nuclei are most tightly bound. This suggests we can "gain"
(that is, release) energy in two ways: below A = 60, by assembling lighter nuclei into heavier nuclei, or above A = 60, by
breaking heavier nuclei into lighter nuclei. In either case we "climb the curve of binding energy" and liberate nuclear energy;
the first method is known as nuclear fusion and the second as nuclear fission.
NUCLEAR BINDING ENERGY
Attempting to understand this curve of binding energy leads us to the semi-empirical formula, in which we try to use a few
general parameters to characterize the variation of with A.

1) The most obvious term to include in estimating is the constant term, since to lowest order . The contribution to the binding
energy from this "volume" term is thus where c1 is a constant to be determined, which should be of order 8 MeV. This linear
dependence of on A is in fact somewhat surprising, and gives us our first insight into the properties of the nuclear force. If every
nucleon attracted all of the others, then the binding energy would be proportional to A(A - 1), or roughly to A2 Since varies
linearly with A, this suggests that each nucleon attracts only its closest neighbors, and not all of the other nucleons. From
electron scattering we learned that the nuclear density is roughly constant, and thus each nucleon has about the same number of
neighbors; each nucleon thus contributes roughly the same amount to the binding energy.

2) An exception to the above argument is a nucleon on the nuclear surface, which is surrounded by fewer neighbors and
thus less tightly bound than those in the central region. These nucleons do not contribute to quite as much as those in
the center, and thus overestimatesby giving full weight to the surface nucleons. We must therefore subtract from a term
proportional to the nuclear surface area. The surface area of the nucleus is proportional to R2 or to A2/3 since V~A and
R~A1/3 (R is the radius of nucleus). Thus the surface nucleons contribute to the binding energy a term of the form •

3) Our binding energy formula must also include the Coulomb repulsion of the protons, which likewise tends to make the
nucleus less tightly bound. Since each proton repels all of the others, this term is proportional to Z( Z - 1), and we may do an
exact calculation, assuming a uniformly charged sphere, to obtain 3 where the negative sign implies a reduction in binding
energy.
NUCLEAR BINDING ENERGY
We also note, from our study of the distribution of stable and radioactive isotopes (Fig.), that stable nuclei have Z≈A/2.

4) If our binding energy formula is to be realistic in describing the stable nuclei that are actually observed, it must take this
effect into account. (Otherwise it would allow stable isotopes of hydrogen with hundreds of neutrons!) This term is very
important for light nuclei, for which Z=A/2 is more strictly observed. For heavy nuclei, this term becomes less important,
because the rapid increase in the Coulomb repulsion term requires additional neutrons for nuclear stability. A possible form for
this term, called the symmetry term because it tends to make the nucleus symmetric in protons and neutrons, is –c4(A - 2Z )2/ A
which has the correct form of favoring nuclei with Z = A/2 and reducing in impor­tance for large A.

5) Finally, we must include another term that accounts for the tendency of like nucleons to couple pairwise to especially stable
configurations. When we have an odd number of nucleons (odd Z and even N, or even Z and odd N ), this term does not
contribute. However, when both Z and N are odd, we gain binding energy by converting one of the odd protons into a neutron
(or vice versa) so that it can now form a pair with its formerly odd partner. We find evidence for this pairing force simply by
looking at the stable nuclei found in nature- there are only four nuclei with odd N and Z (2H, 6Li, 10 B, 14 N), but 167 with even
N and Z. This pairing energy δ is usually expressed as +a5A-3/4 for Z and N even, -a5A-3/4 for Z and N odd, and zero for A odd.

Combining these five terms we get the complete binding energy:

{

𝟏 + a 5 A −3/4    , (𝒆𝒗𝒆𝒏 , 𝒆𝒗𝒆𝒏)
❑ 𝟐 /𝟑 𝟑 𝟐
𝑬 𝑩 =𝐜𝟏 𝐀 − 𝐜𝟐 𝐀 − 𝐜𝟑 𝐙 ( 𝐙− 𝟏 ) 𝐀   – c 4 (A  − 2Z) / A +𝛅 , 𝜹= 𝟎 ,  ( 𝒆𝒗𝒆𝒏 , 𝒐𝒅𝒅 ) 𝒐𝒓 (𝒐𝒅𝒅 , 𝒆𝒗𝒆𝒏)
− a 5 A −3/4 , (𝒐𝒅𝒅 , 𝒐𝒅𝒅)
NUCLEAR BINDING ENERGY
We also note, from our study of the distribution of stable and radioactive isotopes (Fig.), that stable nuclei have Z≈A/2.

4) If our binding energy formula is to be realistic in describing the stable nuclei that are actually observed, it must take this
effect into account. (Otherwise it would allow stable isotopes of hydrogen with hundreds of neutrons!) This term is very
important for light nuclei, for which Z=A/2 is more strictly observed. For heavy nuclei, this term becomes less important,
because the rapid increase in the Coulomb repulsion term requires additional neutrons for nuclear stability. A possible form for
this term, called the symmetry term because it tends to make the nucleus symmetric in protons and neutrons, is –c4(A - 2Z )2/ A
which has the correct form of favoring nuclei with Z = A/2 and reducing in impor­tance for large A.

5) Finally, we must include another term that accounts for the tendency of like nucleons to couple pairwise to especially stable
configurations. When we have an odd number of nucleons (odd Z and even N, or even Z and odd N ), this term does not
contribute. However, when both Z and N are odd, we gain binding energy by converting one of the odd protons into a neutron
(or vice versa) so that it can now form a pair with its formerly odd partner. We find evidence for this pairing force simply by
looking at the stable nuclei found in nature- there are only four nuclei with odd N and Z (2H, 6Li, 10 B, 14 N), but 167 with even
N and Z. This pairing energy δ is usually expressed as +a5A-3/4 for Z and N even, -a5A-3/4 for Z and N odd, and zero for A odd.

Combining these five terms we get the complete binding energy:

{

𝟏 + a 5 A −3/4    , (𝒆𝒗𝒆𝒏 , 𝒆𝒗𝒆𝒏)
❑ 𝟐 /𝟑 𝟑 𝟐
𝑬 𝑩 =𝐜𝟏 𝐀 − 𝐜𝟐 𝐀 − 𝐜𝟑 𝐙 ( 𝐙− 𝟏 ) 𝐀   – c 4 (A  − 2Z) / A +𝛅 , 𝜹= 𝟎 ,  ( 𝒆𝒗𝒆𝒏 , 𝒐𝒅𝒅 ) 𝒐𝒓 (𝒐𝒅𝒅 , 𝒆𝒗𝒆𝒏)
− a 5 A −3/4 , (𝒐𝒅𝒅 , 𝒐𝒅𝒅)
NUCLEAR BINDING ENERGY
Using this expression for we have the semi-empirical mass formula:
𝑀 ( 𝑍 , 𝐴)= 𝑍 𝑚 𝑝 + 𝑁 𝑚𝑛 − 𝐸 𝐵 ( 𝑍 , 𝐴 ) / 𝑐 2

The constants must be adjusted to give the best agreement with the experimental curve. We will use : (  Alonso, Marcelo; Finn,
Edward J. (1969). Fundamental University Physics. Vol. III. Quantum and Statistical Physics. Addison-Wesley Publishing Company. p. 297.)

c1=15.76, c2=17.81, c3=0.711, c4=23.702, c5=34, (in Mev).

In some versions of semi-empirical formula the Coulomb term instead of is presented in form . Thus we have:

{

𝟏 + a 5 A −3/4    , (𝒆𝒗𝒆𝒏 , 𝒆𝒗𝒆𝒏)
❑ 𝟐 /𝟑 𝟐 𝟐
𝑬 =𝐜𝟏 𝐀 − 𝐜𝟐 𝐀
𝑩 − 𝐜𝟑 𝒁 𝐀 𝟑
  – c 4(A  − 2 Z) /  A+𝛅 , 𝛅= 𝟎 ,  ( 𝒆𝒗𝒆𝒏 , 𝒐𝒅𝒅 ) 𝒐𝒓 (𝒐𝒅𝒅 , 𝒆𝒗𝒆𝒏)
− a 5 A −3/4 , (𝒐𝒅𝒅 ,𝒐𝒅𝒅 )
With: c1=15.75, c2=17.8, c3=0.71, c4=23.7, c5=34, (in Mev).

The accuracy of the semi-empirical formula is several MeV. It is not useful for A<20. The equation for binding energy is
precise but it does not represents the intrinsic properties of the nucleus.

One of the first models which could describe very well the behavior of the nuclear binding energies and therefore of nuclear masses was the mass
formula of von Weizsaecker (also called the semi-empirical mass formula – SEMF), that was published in 1935 by German physicist Carl
Friedrich von Weizsäcker. This theory is based on the liquid drop model proposed by George Gamow.
NUCLEAR BINDING ENERGY
The importance of the semi-empirical formula is not that it allows us to predict any new or exotic phenomena of nuclear physics.
Rather, it should be regarded as a first attempt to apply nuclear models to understand the systematic behavior of a nuclear
property, in this case the binding energy. It includes several different varieties of nuclear models: the liquid-drop model, which
treats some of the gross collective features of nuclei in a way similar to the calculation of the properties of a droplet of liquid
(indeed, the first three terms would also appear in a calculation of the energy of a charged liquid droplet), and the shell model,
which deals more with individual nucleons and is responsible for the last two terms of equation.

For constant A, semi-empirical formula represents a parabola of vs Z. The parabola will be centered about the point where
equation reaches a maximum. To compare this result with the behavior of actual nuclei, we must find the maximum, where . We
have (for odd A):
,
where

For small A, Zmax ≈ A/2 as expected, but for large A, Zmax < A/2. For heavy nuclei,
semi-empirical equation gives Z/ A ≈ 0.41, consistent with observed values for
heavy stable nuclei.

We also can to present that dependence of binding energy by plotting mass dependence on Z (accordingly to ) the maximum of
corresponds to minimum of of M:
NUCLEAR BINDING ENERGY

Figure shows a typical odd-A decay chain for A = 125,


leading to the stable nucleus at Z = 52. The unstable nuclei
approach stability by converting a neutron into a proton or a
proton into a neutron by radioactive β decay. Notice how the
decay energy (that is, the mass difference between
neighboring isobars) increases as we go further from
stability.

For even A, the pairing term gives two parabolas, displaced


by 2δ. This permits two unusual effects, not seen in odd-A
decays: (1) some odd-Z, odd-N nuclei can decay in either
direction, converting a neutron to a proton or a proton to a
neutron; (2) certain double β decays can become
energetically possible, in which the decay may change 2 Fig. Mass chains for A = 125 and A = 128. For A = 125, note how
protons to 2 neutrons. Both of these effects are discussed the energy differences between neighboring isotopes increase as
later. we go further from the stable member at the energy minimum. For
A = 128, note the effect of the pairing term; in particular, 128I can
decay in either direction, and it is energetically possible for 128Te to
decay directly to 128 Xe by the process known as double β decay.
NUCLEAR RADIUS
Like the radius of an atom, the radius of a nucleus is not a precisely defined quantity; neither atoms nor nuclei are solid spheres
with abrupt boundaries.
··

The radius that we measure depends on the kind of experiment we are doing to measure the nuclear shape. In some experiments,
such as high-energy electron scattering, muonic X rays, optical and X-ray isotope shifts, and energy differences of mirror nuclei,
we measure the Coulomb interaction of a charged particle with the nucleus. These experiments would then determine the
distribution of protons but also involving somewhat the distribution of neutrons, because of their internal structure). In such
case we have charge radius of nucleus.

In other experiments, such as Rutherford scattering, α-decay, and pionic X rays, we measure the strong nuclear interaction
(strong force interaction between quarks composing nucleons) of nuclear particles, and we would determine the distribution of
nucleons, called the distribution of nuclear matter. In such case we have matter radius of nucleus.

All methods used to define the size of nucleus gives:


𝑅=𝑟 0 𝐴 1/ 3 , 𝑤𝑖𝑡h 𝑟 0=( 1.2 ⩪ 1.4 ) 10 −15 𝑚

We will discuss some of these methods.


NUCLEAR RADIUS: The Distribution of Nuclear Charge
The size of the nucleus could be estimated using the semi-empirical formula. The third term in that formula describes the
reduction of the binding energy due to Coulomb repulsion of the protons. Considering the nucleus as a positively charged
spherical particle with constant density of charge and radius R, from the classical electrodynamics we have Coulomb potential
energy: . Thus we obtain:

And finally, for the radius we have:

1/ 3 3 𝑒2
𝑅=𝑟 𝐴 , 𝑤𝑖𝑡h 𝑟 0=
By taking empirical value =0.71 Mev, we have 0 20 𝜋 𝜀0 𝒄𝟑

The coefficient can be estimated by comparing binding energies of mirror nuclei: the Z of the first nucleus must equal the N of
the second (and thus the N of the first equals the Z of the second). Such pairs of nuclei are called mirror nuclei because one is
changed into the other by "reflecting" in a mirror that exchanges protons and neutrons. Examples of such pairs of mirror nuclei
are and , or and .

If for 1-st) nucleus numbers are Z1 and N1=A-Z1, thus for second we have Z2=Z1-1 and N2=A-Z2 = A-Z1+1. Since for mirror
nuclei N2 = Z1, thus A-Z1+1=Z1 or we have A= 2Z1 -1. So A is the odd number and the last term in semi-empirical formula is
equal to zero. The forth term proportional to (A-2Z)2 = (N-Z)2 = (±1)2 is the same for both nuclei, just the third term in semi-
empirical formula differs.
NUCLEAR RADIUS: The Distribution of Nuclear Charge
For binding energy difference of the mirror nuclei, we obain:
𝟏 𝟏 𝟏 𝟏 𝟏
❑ 𝟐 − 𝟐 − 𝟐 − 𝟐 − 𝟐 −
∆ 𝑬 =− 𝐜𝟑 ( 𝒁 𝟐 ) 𝐀
𝑩
𝟑
+𝐜𝟑 ( 𝒁 𝟏) 𝐀 𝟑
=− 𝐜𝟑 ( 𝒁 𝟏 −𝟏 ) 𝐀 𝟑
+𝐜𝟑 ( 𝒁 𝟏 ) 𝐀 𝟑
=𝐜𝟑 ( 𝟐 𝒁 𝟏 −𝟏 ) 𝐀 𝟑

Since A= 2Z1 -1 , we have: .

Thus, if we know for given A we can find and finally .

For =0.71 Mev, we have

Figure. Coulomb energy differences of mirror nuclei.


The data show the expected A2/3 dependence, and the
slope of the line gives .
NUCLEAR RADIUS: The Distribution of Nuclear Charge
Our usual means for determining the size and shape of an object is to
examine the radiation scattered from it (which is, after all, what we do when
we look at an object or take its photograph). To see the object and its details,
the wavelength of the radiation must be smaller than the dimensions of the
object; otherwise the effects of diffraction will partially or completely
obscure the image.
For nuclei, with a diameter of about 10 fm, we require λ ≤ 10 fm,
corresponding to p ≥ 100 MeV/c. Beams of electrons with energies 100 MeV
to 1 GeV can be produced with high-energy accelerators, and can be analyzed
with a precise spectrometer to select only those electrons that are elastically
scattered from the selected nuclear target. Figure (a) shows an example of
the results of such an experiment.
The first minimum in the diffractionlike pattern can clearly be seen; for
diffraction by a circular disk of diameter D, the first minimum should appear
at θ = sin-1 (1.22 λ/D ), and the resulting estimates for the nuclear radii are
2.6fm for 16O and 2.3fm for 12C. These are, however, only rough estimates
because potential scattering is a three-dimensional problem only
approximately related to diffraction by a two-dimensional disk.

Figure (a) Electron scattering from 16O and 12C. The shape of the cross section is
somewhat similar to that of diffraction patterns obtained with light waves. The
data come from early experiments at the Stanford Linear Accelerator Center (H. F.
Ehrenberg et al., Phys. Rev.113, 666 (1959)).
NUCLEAR RADIUS: The Distribution of Nuclear Charge
Figure (b) also shows how diffuse the nuclear surface appears to be. The charge density is roughly constant out to a certain
point and then drops relatively slowly to zero. The distance over which this drop occurs is nearly independent of the size of the
nucleus, and is usually taken to be constant. We define the skin thickness parameter t as the distance over which the charge
density falls from 90% of its central value to 10%. The value of t is approximately 2.3 fm.

Figure (b). The radial charge distribution of several


nuclei determined from electron scattering. The skin
Figure (c). The rms nuclear radius determined from electron
thickness t is shown for O, Ni, and Pb; its value is
scattering experi­ments. The slope of the straight line gives r0 = 1.23
roughly constant at 2.3 fm. The central density changes
fm. (The line is not a true fit to the data points, but is forced to go
very little from the lightest nuclei to the heaviest. These
through the origin to satisfy the equation R = r0 A1/3 .) The error bars are
distributions were adapted from R. C. Barrett and D. F.
Jackson, Nuclear Sizes and Structure (Oxford: typically smaller than the size of the points ( ± 0.01 fm). More complete
Clarendon, 1977), which gives more detail on methods listings of data and references can be found in the review of C. W. de
of determining p(r). Jager et al., Atomic Data and Nuclear Data Tables 14, 479 (1974).
NUCLEAR RADIUS: The Distribution of Nuclear Matter
An experiment that involves the nuclear force between two nuclei will often provide a measure of the nuclear radius. The
determination of the spatial variation of the force between nuclei enables the calculation of the nuclear radii. In this case the
radius is characteristic of the nuclear, rather than the Coulomb, force; these radii therefore reflect the distribution of all nucleons
in a nucleus, not only the proton
As an example of a measurement that determines the size of the nuclear matter
distribution, we consider an experiment in which a 4He nucleus (α particle) is
scattered from a much heavier target of 197Au. If the separation between the two
nuclei is always greater than the sum of their radii, each is always beyond the
range of the other's nuclear force, so only the Coulomb force acts. (This
situation is known as Rutherford scattering and is discussed in Atom Physics)
The probabil­ity for scattering at a certain angle depends on the energy of the
incident particle exactly as predicted by the Rutherford formula, when the
energy of the incident particle is below a certain value. As the energy of the
incident α-particle is increased, the Coulomb repulsion of the nuclei is
overcome and they may approach close enough to allow the nuclear force to
act. In this case the Rutherford formula no longer holds. Figure shows an
example of this effect.
Figure. The breakdown of the Rutherford scattering formula. When the incident α particle gets
close enough to the target Pb nucleus so that they can interact through the nuclear force (in
addition to the Coulomb force that acts when they are far apart) the Rutherford formula no
longer holds. The point at which this breakdown occurs gives a measure of the size of the
nucleus. Adapted from a review of a particle scattering by R. M. Eisberg and C. E. Porter, Rev.
Mod. Phys. 33, 190 (1961).
NUCLEAR RADIUS: The Distribution of Nuclear Matter

For another example, we consider the form of radioactive decay in which an


α-particle is emitted from the nucleus. The α-particle must escape the
nuclear potential and penetrate a Coulomb potential barrier, as indicated in
Figure. The α decay probabilities can be calculated from a standard barrier-
penetration approach using the Schrodinger equation. These calculated
values depend on the nuclear matter radius R, and comparisons with
measured decay probabilities permit values of R to be deduced (see α-
decay theory below).

The charge and matter radii of nuclei Figure. Barrier penetration in α-decay. The half-life for α-
emission depends on the probability to penetrate the barrier,
are nearly equal, to within about 0.2 fm. which in turn depends on its thickness. The measured half-
lives can thus be used to determine the radius R where the
nuclear force ends and the Coulomb repulsion begins.
𝑅=𝑟 0 𝐴 1/ 3 , 𝑤𝑖𝑡h 𝑟 0=( 1.2 ⩪ 1.4 ) 10 −15 𝑚
NUCLEAR ANGULAR MOMENTUM AND MAGNETIC MOMENTUM
In Section of the Atom Physics we discussed the coupling of orbital angular momentum l and spin s to give total angular
momentum j. To the extent that the nuclear potential is central, l and s (and therefore j) will be constants of the motion. In the
quantum mechanical sense, we can therefore label every nucleon with the corresponding quantum numbers l and s and j. The
total angular momentum of a nucleus containing A nucleons would then be the vector sum of the angular momenta of all the
nucleons.
This total angular momentum is usually called the nuclear spin and is represented by the symbol I. The angular momentum I
has all of the usual properties of quantum mechanical angular momentum vectors: |I|2= ħ21(1 + 1) and Iz = mħ (m = -I, ... , +I).
For many applications involving angular momentum, the nucleus behaves as if it were a single entity with an intrinsic angular
momentum of I.
In ordinary magnetic fields, for example, we can observe the nuclear Zeeman effect, as the state I splits up into its 2I + l
individual substates m = -I, -I + 1, ... , I - 1, I. These sub-states are equally spaced, as in the atomic normal Zeeman effect. If we
could apply an incredibly strong magnetic field, so strong that the coupling between the nucleons were broken, we would see
each individual j splitting into its 2j + l sub-states.
Atomic physics also has an analogy here: when we apply large magnetic fields we can break the coupling between the electronic
l and s and separate the 2l+ 1 components of l and the 2s + 1 components of s. No fields of sufficient strength to break the
coupling of the nucleons can be produced. We therefore observe the behavior of I as if the nucleus were only a single
"spinning" particle. For this reason, the spin (total angular momentum) I and the corresponding spin quantum number I are used
to describe nuclear states.
NUCLEAR ANGULAR MOMENTUM AND MAGNETIC MOMENTUM
To avoid confusion, we will always use I to denote the nuclear spin; we will use j to represent the total angular momentum of
a single nucleon. In nuclei, the pairing of the of nucleons occurs so that their orbital angular momentum and spin angular
momentum each add to zero.
Thus the paired nucleons do not contribute to the magnetic moment, and we need only consider a few valence (outermost)
nucleons. It will often be the case that a single valence particle determines all of the nuclear properties; in that case, I = j. In
other cases, it may be necessary to consider two valence particles, in which case I = j1 + j2 , and several different resultant
values of I may be possible.
One important restriction on the allowed values of I comes from considering the possible z components of the total angular
momentum of the individual nucleons. Each j must be half-integral ( …) and thus its only possible z components are likewise
half-integral ( ).
If we have an even number of nucleons, there will be an even number of half-integral components, with the result that the z
component of the total I can take only integral values. This requires that I itself be an integer.
If the number of nucleons is odd, the total z component must be half-integral and so must the total I. We therefore require the
following rules:
odd-A nuclei: I = half-integer even-A nuclei: I= integer
The measured values of the nuclear spin can tell us a great deal about the nuclear structure. For example, of the hundreds
of known (stable and radioactive) even-Z, even-N nuclei, all have spin-0 ground states. This is evidence for the nuclear
pairing force we discussed in the previous section; the nucleons couple together in spin-0 pairs, giving a total I of zero.
However, the ground state spin of an odd-A nucleus must be equal to the j of the odd proton or neutron.
NUCLEAR ANGULAR MOMENTUM AND MAGNETIC MOMENTUM
The minimum (nonzero) value of number I is I .
All stable odd-odd nuclei (just 5 are existing: ) have non-zero quantum numbers I: (tungsten, volframas), for others I =1. For
all known stable nuclei , except .
Obviously the angular momentum of the nuclei is small in comparison with the sum of absolute values of all nucleons of the
nucleus. This is due to of the pairing of the of nucleons in closed shells so that their orbital angular momentum and spin
angular momentum each add to zero. Thus the paired nucleons do not contribute to the magnetic moment, and we need only
consider a few valence (outermost) nucleons.
The charged particle possesses a magnetic dipole moment µ related to the angular momentum L. For orbital motion (see part
of Atom Physics): . The quantity is called a magneton. For atomic motion we use the electron mass and obtain the Bohr
magneton µB = 5.7884 × 10-5 eV/T.
Putting in the proton mass we have the nuclear magneton µN = 3.1525 × 10-8 eV/T. Note that µN <<µB owing to the difference
in the masses; thus under most circumstances atomic magnetism has much larger effects than nuclear magnetism. For the
quantum mechanical observable (usually projection to z-axis) we have
As we have , thus for maximum observable value (=l) we obtain . For nucleon we can rewrite as: . where gl is the g factor
associated with the orbital angular momentum l. For protons gl= l; because neutrons have no electric charge so to describe the
orbital motion of neutrons if we put gl= 0. The proton magnetic moment in ground state (l=0) is also zero.
NUCLEAR ANGULAR MOMENTUM AND MAGNETIC MOMENTUM
But protons and neutrons, like electrons, also have intrinsic (spin) magnetic mo­ments, which have no classical analog but
which we write in the same form as where s = ½ for protons, neutrons, and electrons. The quantity gs is known as the spin g
factor and is calculated by solving a relativistic quantum mechanical equation.
For a spin- ½ point particle such as the electron gs ≈ 2.0023 ≈ 2. On the other hand, for free nucleons, the experimental values
are far from the expected value for point particles:
proton: gs = 5 .5856912 ± 0 .0000022 neutron: gs= - 3.8260837 ± 0.0000018

Not only is the proton value far from the expected value of 2 for a point particle, but the uncharged neutron has a
nonzero magnetic moment! Here is perhaps our first evidence that the nucleons are not elementary point particles
like the electron, but have an internal structure.

The total magnetic dipole moment of the nucleus µ related to the total angular momentum is usually called the
nuclear spin and is represented by I : where is called nuclear gyromagnetic coefficient.

Sample Values of Nuclear Magnetic Dipole Moments


All values refer to the nuclear ground states; uncertainties are typically
a few parts in the last digit. For a complete tabulation, see V. S. Shirley,
in Table of Isotopes (Wiley: New York, 1978), Appendix VII.
Radioactivity
In 1896 Henri Becquerel was using naturally fluorescent minerals to study the
properties of x-rays, which had been discovered in 1895 by Wilhelm Roentgen.
He exposed potassium uranyl sulfate to sunlight and then placed it on
photographic plates wrapped in black paper, believing that the uranium absorbed
the sun’s energy and then emitted it as x-rays.

This hypothesis was disproved on the 26th-27th of February, when his experiment
"failed" because it was overcast in Paris. For some reason, Becquerel decided to
develop his photographic plates anyway. To his surprise, the images were strong
and clear, proving that the uranium emitted radiation without an external source
of energy such as the sun. Becquerel had discovered radioactivity. It is one of
the most well-known accidental discoveries in the history of physics.

The radioactive decays of naturally occurring minerals containing uranium and


thorium are in large part responsible for the birth of the study of nuclear Henri Becquerel
physics. These decays have half-lives that are of the order of the age of the
Earth, suggesting that the materials are survivors of an early period in the
creation of matter by aggregation of nucleons; the shorter-lived nuclei have
long since decayed away, and we observe today the remaining long-lived
decays.
Radioactivity

In addition to this naturally occurring radioactivity, we also have the capability to


produce radioactive nuclei in the laboratory through nuclear reactions. This was
first done in 1934 by Irene Curie and Pierre Joliot, who used α-particles from the
natural radioactive decay of polonium to bombard aluminum, thereby producing
the isotope 30P (Phosphorus), which they observed to decay through positron
emission with a half-life of 2.5 min.

For this work on artificially produced radioactivity the Joliot-Curie team was
awarded the 1935 Nobel Prize in Chemistry (following a family tradition-Irene's
parents, Pierre and Marie Curie, shared with Becquerel the 1903 Nobel Prize in
Physics for their work on the natural radioactivity of the element radium, and Marie
Curie became the first person twice honored, when she was awarded the 1911 Nobel
Prize in Chemistry).

Marie Curie
THE RADIOACTIVE DECAY
Radioactivity is a spontaneous change of the nucleus composition, which occurs in time period much longer than τ=10 -21 s
(time needed for α-particle with the velocity 10 7 m/s to cross the diameter of typical nucleus 2R~ 10 -14 m ).
Those nuclei that possesses radioactivity property are called radioactive. It is an agreement to call the nucleus as an radioactive
if the radioactive transitions happens in the time interval 10-20 s <T<1010 years. If T>1010 years the nucleus supposed to be
stable.

TYPES of radioactivity:
1. α-decay,
2. β-decay,
3. γ-decay,
4. Spontaneous Fission,
5. Nucleon Emission,
6. Cluster decay, also named heavy particle radioactivity or heavy ion radioactivity
THE RADIOACTIVE DECAY
Alpha decay (or α-decay and also alpha radioactivity) represents the disintegration of a parent nucleus to a daughter
through the emission of the nucleus of a helium atom. This transition can be characterized as:

In this process, a nucleus emits an α-particle (which Rutherford and his co-workers showed to be a nucleus of helium) .
The nucleus shifts by 2 positions to the left in periodic table of elements.

Notice that the number of protons and the number of neutrons must separately be conserved in the decay process. An
example of an α-decay process is

in which the half-life is 1602 years and the α-particle appears with a kinetic energy of about 4.8 MeV.
THE RADIOACTIVE DECAY
Beta decay or β-decay represents the disintegration of a parent nucleus to a daughter through the emission of the beta
particle. This process can occur in three possible ways, each of which must involve another charged particle to conserve
electric charge (the charged particle, originally called a β-particle, was later shown to be identical with ordinary electrons).

1. β- decay (Negative Beta Decay a neutron-rich nucleus emits a high-energy electron (β – particle)):

, mass number A do not change, atomic number Z increase by 1.

2. β+ decay (Positive Beta Decay a proton-rich nucleus emits a positron (positrons are antiparticles of electrons)

, mass number A do not change, atomic number Z decrease by 1.

3. Electron capture (Electron capture, known also as inverse beta decay is sometimes included as a type of beta decay,
because the basic nuclear process, mediated by the weak interaction, is the same.)

, mass number A do not change, atomic number Z decrease by 1.


THE RADIOACTIVE DECAY
The first process is known as negative β-decay and involves the creation and emission of an ordinary electron. The second
process is positive β decay or positron decay, in which a positively charged anti-electron (positron) is emitted. In the third
process, an atomic electron that stays too close to the nucleus is swallowed, allowing the conversion of a proton to a neutron.
In all three processes, yet another particle called a neutrino (or anti-neutrino ) is also emitted, but since the neutrino has no
electric charge, its inclusion in the decay process does not affect the identity of the other final particles.

In contrast to alpha decay, neither the beta particle nor its associated neutrino exist within the nucleus prior to beta decay, but
are created in the decay process. By this process, unstable atoms obtain a more stable ratio of protons to neutrons. The
probability of a nuclide decaying due to beta and other forms of decay is determined by its nuclear binding energy. For either
electron or positron emission to be energetically possible, the energy release must be positive.

Beta decay is governed by the weak interaction. During beta decay one of two down quarks (q=-1/3e) changes into an up quark
(q=+2/3e) by emitting a W– boson (carries away a negative charge). The W– boson then decays into a beta particle and
an antineutrino. This process is equivalent to the process, in which a neutrino interacts with a neutron.
THE RADIOACTIVE DECAY
Gamma decay or γ decay represents the disintegration of a parent nucleus to a daughter through the emission of gamma
rays (high energy photons). This transition (γ decay) can be characterized as:

As can be seen, if a nucleus emits a gamma ray, atomic and mass numbers of daughter nucleus remain the same, but daughter
nucleus will form different energy state of the same element. Nuclides with equal proton number and equal mass number
(thus making them by definition the same isotope), but in a different energy state are known as nuclear isomers. We usually
indicate isomers with a superscript m, thus: 241mAm or 110mAg.

In most practical laboratory sources, the excited nuclear states are created in
the decay of a parent radionuclide, therefore a gamma decay
typically accompanies other forms of decay, such as alpha or beta decay.
Typically after a beta decay (isobaric transition), nuclei usually contain too
much energy to be in its final stable daughter state.

Gamma rays are high-energy photons with very short wavelengths and thus


very high frequency. Gamma rays from radioactive decay are in the energy
range from a few keV to ~8 MeV, corresponding to the typical energy levels in
nuclei with reasonably long lifetimes.
THE RADIOACTIVE DECAY
In contrast to alpha and beta radioactivity, gamma radioactivity is governed by an electromagnetic interaction rather than
a weak or strong interaction. As in atomic transitions, the photon carries away at least one unit of angular momentum (the
photon, being described by the vector electromagnetic field, has spin angular momentum of ħ.Gamma decay may
follow nuclear reactions such as neutron capture, nuclear fusion, or nuclear fission.

Most of nuclear reactions produce extremely unstable nuclei that decay as soon as they are formed in nuclear reactions
(half-life less than 10-11s) and are not generally classified as nuclear isomers. Moreover, these nuclei usually produce
a cascade of gamma rays and the gamma-ray cascade concludes when all excess energy of the excited nucleus is released.

In certain cases, the excited nuclear state that follows the emission of a beta particle or other type of excitation, are able
to stay in metastable state for a long time (hours, days and sometimes much longer) before undergoing gamma decay, in
which they emit a gamma ray. These long-lived excited nuclei are known as isomeric states (or isomers) and their decays are
termed isomeric transitions. The process of isomeric transition is therefore similar to any gamma emission, but differs in
that it involves the intermediate metastable excited state(s) of the nuclei.

Metastable nuclei are often characterized by high nuclear spin, requiring a change in spin of several units or more with
gamma decay, instead of a single unit transition that occurs in only 10 −12 seconds. The rate of gamma decay is also slowed
when the energy of excitation of the nucleus is small. An example is the
decay of the isomer or metastable state of protactinium:
Extremely unstable nuclei that decay as soon as they are formed in nuclear reactions
(half-life less than 10-11s) are not generally classified as nuclear isomers. Isomeric
transitions must occur by higher order multipole transitions (in contrast to gamma
emission that occurs by dipole radiation) that occur on a longer time-scale.
THE RADIOACTIVE DECAY
Spontaneous Fission
The occurrence of the spontaneous fission of a heavy nucleus was first observed by Petrzhak and Flerov (1940), who
detected the spontaneous fission of natural uranium. In this process the nucleus splits into two fragment nuclei of roughly
half the mass of the parent. This process is only barely detectable in competition with the more prevalent alpha
decay for uranium, but for some of the heaviest artificial nuclei, such as fermium-256, spontaneous fission becomes the
predominant mode of radioactive decay.

Proton decay is a rare type of radioactive decay of nuclei containing excess protons, in which a proton is simply ejected from
the nucleus. This article describes mainly spontaneous proton emission (proton decay) and does not describe decay of a
free proton. Note that, a free proton (a proton not bound to nucleons or electrons) is a stable particle that has not been
observed to break down spontaneously to other particles. Proton radioactivity, discovered in 1970, is exhibited by an excited
isomeric state of cobalt-53, 53mCo, 1.5 percent of which emits protons:

Spontaneous neutron emission. Spontaneous neutron emission is a mode of radioactive decay in which one or more
neutrons are ejected from a nucleus.
THE RADIOACTIVE DECAY

Cluster decay, also named heavy particle radioactivity or heavy ion radioactivity, is a rare type of nuclear
decay in which an atomic nucleus emits a small "cluster“ of neutrons and protons, more than in an alpha
particle, but less than a typical binary fission fragment. 

In 1980 A. Sandulescu, D.N. Poenaru, and W. Greiner described calculations indicating the possibility of a new type of
decay of heavy nuclei intermediate between alpha decay and spontaneous fission. The first observation of heavy-ion
radioactivity was that of a 30-MeV, carbon-14 emission from radium-223 by H.J. Rose and G.A. Jones in 1984.

223
88
Ra → 146C + 20982Pb

The ratio of carbon-14 decay to alpha decay is about 5 × 10 −10. Observations also have been made of carbon-14 from
radium-222, radium-224, and radium-226, as well as neon-24 from thorium-230, protactinium-231, and uranium-232.
Such heavy-ion radioactivity, like alpha decay and spontaneous fission, involves quantum-mechanical tunneling through
the potential-energy barrier. Shell effects play a major role in this phenomenon, and in all cases observed to date the
heavy partner of carbon-14 or neon-24 is close to doubly magic lead-208 (see  Nuclear models).
THE RADIOACTIVE SERIES
Radioactive series (known also as a radioactive cascades) are three naturally occurring radioactive decay chains and one
artificial radioactive decay chain of unstable heavy atomic nuclei that decay through a sequence of alpha and beta
decays until a stable nucleus is achieved.

Most radioisotopes do not decay directly to a stable state and all isotopes within the series decay in the same way. In
physics of nuclear decays, the disintegrating nucleus is usually referred to as the parent nucleus and the nucleus
remaining after the event as the daughter nucleus.

Since alpha decay represents the disintegration of a parent nucleus to a daughter through the emission of the nucleus of
a helium atom (which contains four nucleons), there are only four decay series. Within each series, therefore, the mass
number of the members may be expressed as four times an appropriate integer (n) plus the constant for that series. As a
result, the thorium series is known as the 4n+0 series, the neptunium series as the 4n + 1 series, the uranium series as
the 4n + 2 series and the actinium series as the 4n + 3 series.

Three of the sets are called natural or classical series. The fourth set, the neptunium series, is headed by neptunium-237.
Its members are produced artificially by nuclear reactions and do not occur naturally.

the thorium series (4n series),


the uranium series (4n+2 series),
the actinium series (4n+3 series),
the neptunium series (4n+1 series).
THE RADIOACTIVE SERIES
The classical series are headed by primordial unstable nuclei (Bismuth, thorium, uranium and plutonium are primordial
nuclides because they have half-lives long enough to still be found on the Earth, while all the others are produced either
by radioactive decay or are synthesized in laboratories and nuclear reactors.). Primordial nuclides are nuclides found on the
Earth that have existed in their current form since before Earth was formed. The previous four series consist of the
radioisotopes, that are the descendants of four heavy nuclei with long and very long half-lives:

the thorium series with thorium-232 (with a half-life of 14.0 billion years),
the uranium series with uranium-238 (which lives for 4.47 billion years),
the actinium series with uranium-235 (with a half-life of 0.7 billion years).
the neptunium series with neptunium-237 (with a half-life of 2 million years).

The half-lives of all the daughter nuclei are all extremely variable, and it is difficult to represent a range of timescales going
from individual seconds to billions of years. Since daughter radioisotopes have different half-lives then secular equilibrium is
reached after some time. In the long decay chain for a naturally radioactive element, such as uranium-238, where all of the
elements in the chain are in secular equilibrium, each of the descendants has built up to an equilibrium amount and all
decay at the rate set by the original parent. If and when equilibrium is achieved, each successive daughter isotope is present
in direct proportion to its half-life. Since its activity is inversely proportional to its half-life, each nuclide in the decay chain
finally contributes as many individual transformations as the head of the chain.

As can be seen from figures, branching occurs in all four of the radioactive series. That means the decay of a given species
may occur in more than one way. For example, in the thorium series, bismuth-212 decays partially by negative beta
emission to polonium-212 and partially by alpha emission to thallium-206.
THE RADIOACTIVE SERIES
The thorium series is one of three classical
radioactive series beginning with naturally
occurring thorium-232. This radioactive decay
chain consists of unstable heavy atomic nuclei
that decay through a sequence
of alpha and beta decays until a stable nucleus
is achieved. In case of thorium series, the stable
nucleus is lead-208.

Since alpha decay represents the disintegration


of a parent nucleus to a daughter through the
emission of the nucleus of a helium atom
(which contains four nucleons), there are only
four decay series. Within each series, therefore,
the mass number of the members may be
expressed as four times an appropriate integer
(n) plus the constant for that series.
As a result, the thorium series is known as
the 4n series.

The total energy released from thorium-232 to


lead-208, including the energy lost to neutrinos,
is 42.6 MeV.
THE RADIOACTIVE SERIES
The neptunium series is a radioactive series
beginning with neptunium-237. Its members
are produced artificially by nuclear reactions
 and do not occur naturally, because the half-
life of the longest lived isotope in the series is
short compared to the age of the earth. 

This radioactive decay chain consists of unstable


heavy atomic nuclei that decay through a
sequence of alpha and beta decays until a
stable nucleus is achieved. In case of neptunium
series, the stable nucleus is bismuth-209 (with
half-life of 1.9E19 years) and thallium-205.

The neptunium series is known as the 4n+1


series.

The total energy released from neptunium-237 Bismuth-209 was long thought to have the heaviest stable nucleus of any
to thallium-205, including the energy lost element, but in 2003, a research team at the Institut d’Astrophysique Spatiale
to neutrinos, is 50.0 MeV. in Orsay, France, discovered that 209Bi undergoes alpha decay with a half-life of
approximately 1.9×1019years, over a billi`on times longer than the current
estimated age of the universe. Primordial bismuth consists entirely of this isotope.
THE RADIOACTIVE SERIES
The uranium series, known also as radium
series, is one of three classical radioactive series
beginning with naturally occurring uranium-238
. This radioactive decay chain consists of
unstable heavy atomic nuclei
that decay through a sequence
of alpha and beta decays until a stable nucleus
is achieved.

In case of uranium series, the stable nucleus is


lead-206.

Within each series, therefore, the mass number


of the members may be expressed as four times
an appropriate integer (n) plus the constant for
that series.

As a result, the uranium series is known as


the 4n+2 series.

The total energy released from uranium-238 to


lead-206, including the energy lost to neutrinos,
is 51.7 MeV.
THE RADIOACTIVE SERIES
The actinium series is one of three classical
radioactive series beginning with naturally
occurring uranium-235. This radioactive decay
chain consists of unstable heavy atomic nuclei
that decay through a sequence of alpha and 
beta decays until a stable nucleus is achieved. In
case of actinium series, the stable nucleus is
lead-207.

Within each series, therefore, the mass number


of the members may be expressed as four times
an appropriate integer (n) plus the constant for
that series.

As a result, the actinium series is known as


the 4n+3 series.

The total energy released from uranium-235 to


lead-207, including the energy lost to neutrinos,
is 46.4 MeV.
THE RADIOACTIVE DECAY LAW
Three years following the 1896 discovery of radioactivity it was noted that the decay rate of a pure radioactive substance
decreases with time according to an exponential law. It took several more years to realize that radioactivity represents changes
in the individual atoms and not a change in the sample as a whole. It took another two years to realize that the decay is
statistical in nature, that it is impossible to predict when any given atom will disintegrate, and that this hypothesis leads
directly to the exponential law.

If N radioactive nuclei are present at time t and if no new nuclei are introduced into the sample, then the number dN decaying
in a time dt is proportional to N, and so
in which is a constant called the disintegration or decay constant.

Integrating leads to the exponential law of radioactive decay:


where N0 is the original number of nuclei present at t = 0. The half-life gives the time necessary for half of the nuclei to
decay. Putting N = N0 / 2 in Equation gives

It is also useful to consider the mean lifetime (sometimes called just the lifetime) τ which is defined as the average time that a
nucleus is likely to survive before it decays. The number that survive to time t is just N(t), and the number that decay between
t and t+dt is |dN /dt|dt. The mean lifetime is then

Thus the mean lifetime is simply the inverse of the decay constant.
THE RADIOACTIVE DECAY LAW
Defining the activity to be the rate at which decays occur in the sample,

Units of Activity
Becquerel. The becquerel is SI unit of radioactivity defined in 1974. It is named in honour of Henri Becquerel, a French physicist
who discovered radioactivity in 1896. One becquerel (1Bq) is equal to 1 disintegration per second.

Curie. The curie is a non-SI unit of radioactivity defined in 1910. It was originally defined as equivalent to the number of
disintegrations that one gram of radium-226 will undergo in one second. Currently, a curie is defined as 1Ci = 3.7 x
1010 disintegrations per second.

Rutherford. Rutherford (symbol Rd) is also a non-SI unit defined as the activity of a quantity of radioactive material in which one
million nuclei decay per second (1Rd=MBq, one megabecquerel, 2.703 × 10−5 curie.).

Often it will happen that a given initial nucleus can decay in two or more different ways, ending with two different final nuclei.
Let's call these two decay modes a and b. The rate of decay into mode a, (dN/dt)a is determined by the partial decay constant λa,
and the rate of decay into mode b, (dN/dt)b is determined by the partial decay constant λb. The total decay rate is

Where is the total decay constant.


THE RADIOACTIVE DECAY LAW
Let's denote by N1 the number of radioactive nuclei that are formed as a result of the reaction. These nuclei decay with decay
constant λ1 to the stable nuclei denoted by N2. Thus the number of nuclei N1 present increases owing to the production at the rate
Q and decreases owing to the radioactive decay:

and the solution to this equation is easily obtained:

If the irradiation time is short compared with one half-life, then we can expand the exponential and keep only the term linear in
t:

For small times, the activity thus increases at a constant rate. This corresponds to the linear (in time) accumulation of product
nuclei, whose number is not yet seriously depleted by radioactive decays.

For times long compared with the half-life the exponential approaches zero and the activity is approximately constant:

Another common situation occurs when a radioactive decay results in a product nucleus that is also radioactive. It is thus
possible to have series or chains of radioactive decays 1→ 2 → 3 → 4 → ... , and it has become common to refer to the original
nucleus (type 1) as the parent and the succeeding "generations" as daughter (type 2), granddaughter (type 3), and so on.
We assume that we begin with N0 atoms of the parent at t = 0 and that no atoms of the decay products are originally present:

N1(t = 0) = N0 , N2 (t = 0) = N3(t = 0) = 0
THE RADIOACTIVE DECAY LAW

For the present calculation, we will assume that the granddaughter is the stable end-product of the decay. The number of
parent nuclei decreases with time according to the usual form:

The number of daughter nuclei increases as a result of decays of the parent and decreases as a result of its own decay:

The number of parent nuclei can be found directly from integrating equation:

To solve ,we try a solution of the form: . Using the initial condition N2(t = 0)=0 we find

Note that equation reduces to equation if nuclei of type 2 are stable


(𝜆1 ≪ 𝜆2) THE RADIOACTIVE DECAY LAW
Let's suppose that , is very small (but not quite zero), so that: and , thus we have

Thus the activity approaches the limiting value as was shown in Figure

This is an example of secular equilibrium, where as t


becomes large nuclei of type 2 are decaying at the same
rate at which they are formed: . (Note that Equation
shows immediately that dN2 /dt = 0 in this case.
( 𝜆1 < 𝜆2) THE RADIOACTIVE DECAY LAW
We can calculate the ratio of the two activities:

As t increases, the exponential term becomes smaller and the


ratio approaches the limiting constant value .

The activities themselves are not constant, but the nuclei of type
2 (daughter) decay (in effect) with the decay constant of type 1
(parent). This situation is known as transient equilibrium and is
illustrated in Figure.

( 𝜆1 > 𝜆2)
In this case the parent decays quickly, and the daughter activity
rises to a maximum and then decays with its characteristic decay
constant. When this occurs the number of nuclei of type 1 is
small and nearly insignificant. If we have
THE RADIOACTIVE DECAY LAW
If we now assume that there are several succeeding
generations of radioactive nuclei (that is, the
granddaughter nuclei type 3 are themselves radioactive,
as are types 4, 5, 6, ... ), we can then easily generalize
previous equations.

In physics, the Bateman equations are a set of first-order


differential equations, which describe the time evolution
of nuclide concentrations undergoing serial or linear
decay chain. The model was formulated by Ernest
Rutherford in 1905 and the analytical solution for the
case of radioactive decay in a linear chain was provided
by Harry Bateman in 1910. This model can be also used
in nuclear depletion codes to solve nuclear
transmutation and decay problems.

The Bateman equations for radioactive decay case of n –


nuclide series in linear chain describing nuclide
concentrations are as follows:
α-decay
Alpha decay (or α-decay and also alpha radioactivity) represents the disintegration of a parent nucleus to a daughter
through the emission of the nucleus of a helium atom.

Today are known ≈ 200 an α-radioactive nuclei, most of them are created artificially in nuclear reactions. The α-decay
is possible only if the binding energy of parent particle is less than sum of binding energy of daughter particle and α-
particle:

The released kinetic energy is :


.

Since the mass of α-particle is much smaller than the mass of daughter nucleus, thus the most of this kinetic energy is
the kinetic energy of α-particle. The α-decay is possible just when Q > 0. We can represent the equation for Q by using
the binding energy per nucleon : . Thus for Q we have:

Indices “p” and “d” denotes parent and daughter nuclei.


α-decay
Thus we obtain: . For all A>>1 we have (see Fig.)

Thus, when A >> 1 the condition for α-decay Q > 0 can be reformulated as

Thus binding energy per nucleon should decrease. In practice the α-decay
becomes possible starting at Z=60 and A=144, or from isotope:

(Cerium)

One feature of α-decay is so striking that it was noticed as long ago as 1911, the year that Rutherford "discovered" the nucleus.
Geiger and Nuttall noticed that α-emitters with large disintegration energies had short half-lives and conversely. The variation
is astonishingly rapid as we may see from the limiting cases of (1.4 x 10 10 years; Q = 4.08 MeV) and (1.0 x 10-7 s; Q = 9.85
MeV). A factor of 2 in energy means a factor of 1024 in half-life!
These experimenters had found that , for many decays is approximately proportional to where the kinetic energy of the
emitted α-particle:
𝐷 Geiger and Nuttall
𝑙𝑜𝑔 𝑇 1 =𝐶 +
2 √𝐸𝛼 Law
α-decay

𝐷 Geiger and Nuttall


𝑙𝑜𝑔 𝑇 1 =𝐶 +
2 √𝐸𝛼 Law

A plot of log against in which all α-emitters are


included shows a considerable scatter about the general
Geiger-Nuttall trend. Very smooth curves result,
however, if we plot only α-emitters with the same Z and
if further we select from this group only those with Z
and N both even (see Fig.). Even-odd, odd-even, and
odd-odd nuclei obey the general trend but do not plot
into quite such smooth curves; their periods are 2-1000
times longer than those for even-even types with the
same Z and .
Fig. The inverse relationship between α-decay half-life and decay
energy, called the Geiger-Nuttall rule. Only even-Z, even-N nuclei are
shown. The solid lines connect the data points.
α-decay theory
By 1928, George Gamow (and independently by Ronald Gurney and Edward Condon)
had solved the theory of alpha decay via quantum tunneling. They assumed that the
alpha particle and the daughter nucleus exist within the parent nucleus prior to its
dissociation, namely the decay of quasi-stationary states (QS).  

A quasi-stationary state is defined as a long-lived state that eventually decays. Initially,


the alpha cluster oscillates in the potential of the daughter nucleus, with the Coulomb
potential preventing their separation. The alpha particle is trapped in a potential well
by the nucleus (nuclear forces). Classically, it is forbidden to escape, but according to
the principles of quantum mechanics, it has a tiny (but non-zero) probability of
“tunneling” through the barrier and appearing on the other side to escape the
nucleus.

Using the tunneling mechanism, Gamow, Condon and Gurney calculated the
penetrability of the tunneling α particle through the Coulomb barrier, finding the Alpha decay is a quantum
lifetimes of some α emitting nuclei. tunneling process. In order to be
emitted, the alpha particle must
The main success of this model was the reproduction of the semi-empirical Geiger- penetrate a potential barrier.
Nuttall law that expresses the lifetimes of the α emitters in terms of the energies of This is similar to cluster decay, in
the released α particles. It must be noted, that other common forms of decay (e.g. which an atomic nucleus emits a
beta decay) are governed by the interplay between both the nuclear force and the small “cluster” of neutrons and
electromagnetic force. protons (e.g. 12C).
α-decay theory
In the theory of Gamow an α-particle is assumed to move in a spherical region determined
by the daughter nucleus. The central feature of this one-body model is that the α-particle is
preformed inside the parent nucleus. Actually there is not much reason to believe that α-
particles do exist separately within heavy nuclei; nevertheless, the theory works quite well,
especially for even-even nuclei. This success of the theory does not prove that α-particles
are preformed, but merely that they behave as if they were.
R
Figure shows a plot, suitable for purposes of the theory, of the potential energy between the
α-particle and the horizontal line is the disintegration energy. Note that the Coulomb
potential is extended inward to a radius R and then arbitrarily cut off. The radius R can be
taken as the sum of the radius of the residual nucleus and of the α-particle.

There are three regions of interest. In the spherical region r < R we are inside the nucleus
and speak of a potential well of depth - V0 , where V0 is taken as a positive number.
Classically the α-particle can move in this region, with a kinetic energy but it cannot 𝐵
escape from it.

The annular­ shell region R < r < r1 forms a potential barrier because here the potential
energy is more than the residual nucleus for various distances between their centers. The
total available energy . Classically the a particle cannot enter this region from either
direction, just as a tennis ball dropped from a certain height cannot rebound higher. The
region r > r1 is a classically permitted region outside the barrier.
α-decay theory
From the classical point of view, an α-particle in the spherical potential well would
sharply reverse its motion every time it tried to pass beyond r = R. Quantum
mechanically, however, there is a chance of "leakage" or "tunnelling" through such a
barrier. This barrier accounts for the fact that α-unstable nuclei do not decay immediately.
The α-particle within the nucleus must present itself again and again at the barrier surface
until it finally penetrates. In 238U, for example, the leakage probability is so small that the R
α-particle, on the average, must make ~ 10 38 tries before it escapes ( ~ 1021 per second for
~ 109 years)!

The barrier also operates in reverse, in the case of α-particle scattering by nuclei. Alpha
particles incident on the barrier from outside the nucleus usually scatter in the Coulomb
field if the incident energy is well below the barrier height. Tunnelling through the
barrier, so that the nuclear force between the particle and target can cause nuclear
reactions, is a relatively improbable process at low energy. The theoretical analysis of
𝐵
nuclear reactions induced by charged particles uses a formalism similar to that of α-decay
to calculate the barrier penetration probability. Fusion reactions, such as those responsible
for the energy released in stars, also are analyzed using the barrier penetration approach.

The disintegration constant of an α-emitter is given in the one-body theory by

where f is the frequency with which the α-particle presents itself at the barrier and P is the
probability of transmission through the barrier.
α-decay theory
The disintegration constant of an α-emitter is given in the one-body theory by

where f is the frequency with which the α-particle presents itself at the barrier and P is the
probability of transmission through the barrier.
R
The quantity f is roughly of the order of v/R where v is the relative velocity of the α-
particle as it rattles about inside the nucleus. We can find v in to ways:

1-st) from the kinetic energy of the α-particle at r < R, we have

Estimating V0 = 35 MeV for a typical well depth gives f= 6 x 1021/ s for =5 MeV.

2-nd) from uncertainty principle


𝐵

We will see later that we do not need to know f very precisely to check the theory.
α-decay theory
The barrier penetration probability P must be obtained from a quantum mechanical
calculation. The result depends on the width of the barrier and on its height (called V0 for
the rectangular barrier) above the energy of the particle. The Coulomb barrier (see Figure)
has height B at r = R :

The α-particle has charge 2e and the daughter nucleus has charge Ze.
For a typical heavy nucleus (Z = 90, R = 7.5 fm), the barrier height B is about 34 MeV.

The radius r1 at which the α-particle "leaves" the barrier is found from the equality of the
particle's energy and the potential energy:
𝐵

for a typical case of a heavy nucleus with = 6 MeV, = 42 fm.

The probability to penetrate the complete barrier is

where the Gamow factor G is .

Substituting Coloumb potential and integrating, we obtain


α-decay theory
Substituting Coloumb potential and integrating, for we obtain:

For Z=90, R=10-14, B=26MeV, =4.2MeV, we obtain years, the experimental value is years, so
coincidence not so bad.

For the mean lifetime we obtain: If <<B, then , , we have


𝐵
This is just other form of expression of Geiger and Nuttall Law.
α-decay theory
The results of this calculation for the even isotopes of Th are shown in Table. The agreement is not exact, but the calculation is
able to reproduce the trend of the half-lives within 1-2 orders of magnitude over a range of more than 20 orders of magnitude.
Calculated α-Decay Half-lives for Th Isotopes
---------------------------------------------------------------
A (s) (s)
Measured Experimental
222 8.13 2.8 X 10-3 6.3 X 10-5
224 7.31 1.04 3.3 X 10-2
226 6.45 1854 6.0 X 101
228 5.52 6.0 X 107 2.4 X 106
230 4.77 2.5 X 1012 1.0 X 1011
232 4.08 4.4 X 1017 2.6 X 1016
Even though this oversimplified theory is not strictly correct, it gives us a good estimate of the decay half-lives. It also
enables us to understand why other decays into light particles are not commonly seen, even though they may be allowed by
the value. For example, the decay 220Th → 12C + 208Po would have a value of 32.1 MeV, and carrying through the calculation
using Equation presented above gives = 2.3 x 106 s for the 220Th decay into 12C. This is a factor of 1013 longer than the α-
decay half-life and thus the decay will not easily be observable.

Recently, just such a decay mode has in fact been observed, the first example of a spontaneous decay process involving
emission of a particle heavier than an α. The decay of 223Ra normally proceeds by α-emission with a half-life of 11.2 d, but
there has now been discovered the decay process 223Ra→14C + 209Pb. The probability for this process is very small, about 10 -9
relative to the α-decay. The total of 11 events observed represents about six months counting !
Passage of Heavy charged particles through matter
Heavy charged particles are all energetic ions with mass of one atomic mass unit or greater, such as protons, alpha
particles (helium nuclei) or fission fragments. Especially knowledge of the interaction of fission fragments and alpha particles
must be well known in the engineering of nuclear reactors. Heavy charged particles, such as fission fragments or alpha
particles interact with matter primarily through coulomb forces between their positive charge and the negative charge of the
electrons from atomic orbitals. Alpha particles are relatively large and carry a double positive charge. They are not very
penetrating and a piece of paper can stop them. They travel only a few centimeters but deposit all their energies along their
short paths. In nuclear reactors they are produced for example in the fuel (alpha decay of heavy nuclei).
Alpha particles are commonly emitted by all of the heavy radioactive nuclei occuring in the nature (uranium,
thorium or radium), as well as the transuranic elements (neptunium, plutonium or americium). Especially energetic alpha
particles (except artificially accelerated helium nuclei) are produced in a nuclear process, which is known as a ternary fission.
In this process, the nucleus of uranium is splitted into three charged particles (fission fragments) instead of the normal two.
The smallest of the fission fragments most probably (90% probability) being an extra energetic alpha particle.
Since the electromagnetic interaction extends over some distance, it is not necessary for the light or heavy charged particle to
make a direct collision with an atom. They can transfer energy simply by passing close by. Heavy charged particles, such as
fission fragments or alpha particles interact with matter primarily through coulomb forces between their positive charge and
the negative charge of the electrons from atomic orbitals. On the other hand, the internal energy of an atom is quantised,
therefore only certain amount of energy can be transferred. In general, charged particles transfer energy mostly by:

Excitation. The charged particle can transfer energy to the atom, raising electrons to a higher energy levels.

Ionization. Ionization can occur, when the charged particle have enough energy to remove an electron. This results in a
creation of ion pairs in surrounding matter.
Passage of Heavy charged particles through matter
Creation of pairs requires energy, which is lost from the kinetic energy of the charged particle causing it to decelerate. The
positive ions and free electrons created by the passage of the charged particle will then reunite, releasing energy in the form of
heat (e.g. vibrational energy or rotational energy of atoms). This is the principle how fission fragments heat up fuel in the 
reactor core. There are considerable differences in the ways of energy loss and scattering between the passage of light charged
particles such as positrons and electrons and heavy charged particles such as fission fragments, alpha particles, muons. Most of
these differences are based on the different dynamics of the collision process.

In general, when a heavy particle collides with a much lighter particle (electrons in the atomic orbitals), the laws of energy and
momentum conservation predict that only a small fraction of the massive particle’s energy can be transferred to the less
massive particle. The actual amount of transferred energy depends on how closely the charged particles passes through the
atom and it depends also on restrictions from quantisation of energy levels.

The distance required to bring the particle to rest is referred to as its range. The range of fission fragments in solids amounts to
only a few microns, and thus most of the energy of fission is converted to heat very close to the point of fission. In case of
gases the range increases to a few centimeters in dependence of gas parameters (density, type of gas etc.)  The trajectory of
heavy charged particles are not greatly affected, because they interacts with light atomic electrons. Other charged particles,
such as the alpha particles behave similarly with one exception – for lighter charged particles the ranges are somewhat longer.

A convenient variable that describes the ionization properties of surrounding medium is the stopping power. The
linear stopping power of material is defined as the ratio of the differential energy loss for the particle within the material to
the corresponding differential path length:
Passage of Heavy charged particles through matter
The linear stopping power of material is defined as the ratio of the differential energy loss for the particle within the material
to the corresponding differential path length:

where E is the kinetic energy of the charged particle.


For charged particles, S increases as the particle velocity decreases. The classical expression that describes the specific energy
loss is known as the Bethe  formula. The non-relativistic formula was found by Hans Bethe in 1930. The relativistic version (see
below) was found also by  Hans Bethe in 1932.

In this expression, me is the rest mass of the electron, β equals to v/c, v is the particle velocity, z is the atomic number of the
particle, n is the medium atoms density in the volume. I represents the mean excitation energy of the atomic electrons, which
could in principle be computed by averaging over all atomic ionization and excitation processes. In practice, I is regarded
as an empirical constant, with a value in eV of the order of I=13.5xZ. For nonrelativistic particles (heavy charged particles are
mostly nonrelativistic), dE/dx is dependent on 1/v2. This is can be explained by the longer time the charged particle spends in
the negative field of the electron, when the velocity is low.
Passage of Heavy charged particles through matter
The stopping power of most materials is very high for heavy charged particles and these particles have very short ranges. For
example, the range of a 5 MeV alpha particle is approximately only 0,002 cm in aluminium alloy. Most alpha particles can be
stopped by an ordinary sheet of paper or living tissue.

The mass stopping power of a material is obtained by dividing the stopping power by the density ρ.
.
Since for electron density , (𝑍 is the atomic number of medium, 𝜌 is the medium density, 𝑁𝐴 is Avogadro number, is
the atomic weight (moll mass) of medium. ) We have:

Since: . (A is the mass number)


Common units for mass stopping power, -dE/ρdx, are MeV cm2 g-1.

In a gas, for example, -dE/dx depends on pressure, but –dE/ρdx does not, because dividing by the density exactly
compensates for the pressure.

Mass stopping power does not differ greatly for materials with similar atomic composition.

Mass stopping powers for water can be scaled by density and used for tissue, plastics, hydrocarbons, and other materials that
consist primarily of light elements.
Range for Heavy Charged Particles
There are two related definitions of the range of heavy charged particles:

Mean range: the absorber thickness that reduces the alpha particle count
to exactly one-half of its value in the absence of the absorber.

Extrapolated range: extrapolating the linear portion of the end of the


transmission curve to zero.
The mean range of any heavy charged particle can be related to the stopping power as the following: .
From Bethe formula we have:

And from , thus

The function is the same for all particles.


Assuming weak dependence of logarithmic function on v, we have
Passage of Heavy charged particles through matter
The Bragg curve is typical for heavy charged particles and
describes energy loss of ionizing radiation during travel through
matter. For this curve is typical the Bragg peak, which is the result
of 1/v2 dependency of the stopping power. This peak occurs
because the cross section of interaction increases immediately
before the particle come to rest. For most of the track, the charge
remains unchanged and the specific energy loss increases
according to the 1/v2. Near the end of the track, the charge can
be reduced through electron pickup and the curve can fall off.

The Bragg curve also differs somewhat due to the effect of


straggling (besiblaškantis). For a given material the range will be
the nearly the same for all particles of the same kind with
the same initial energy. Because the details of the microscopic As charged particle penetrates matter, statistical
interactions undergone by any specific particle vary randomly, fluctuation occur in the number of collisions along its
a small variation in the range can be observed. This variation is path and in the amount of energy lose in each collision.
called straggling and it is caused by the statistical nature of the As a result, a number of identical particles starting out
energy loss process which consists of a large number of individual under identical conditions will show (1) a distribution of
collisions. energies as they pass a given depth – the energy
This phenomenon, which is described by the Bragg curve, is straggling and (2) a distribution of path-lengths traversed
exploited in particle therapy of cancer, because this allows to before they stop – the range straggling.
concentrate the stopping energy on the tumor while minimizing
the effect on the surrounding healthy
β-decay
Beta decay or β decay represents the disintegration of a parent nucleus to a daughter through the emission of the beta
particle. This process can occur in three possible ways, each of which must involve another charged particle to conserve
electric charge (the charged particle, originally called a β particle, was later shown to be identical with ordinary electrons).

1. β- decay (Negative Beta Decay a neutron-rich nucleus emits a high-energy electron (β – particle)):

, mass number A do not change, atomic number Z increase by 1.

The β- decay is possible when the mass (energy) constraint is fulfilled:

or for atomic masses:

Thus the spontaneous β- decay is possible when the mass of atom becomes less when number Z is increasing.
β-decay
2. β+ decay (Positive Beta Decay a proton-rich nucleus emits a positron (positrons are antiparticles of electrons)

, mass number A do not change, atomic number Z decrease by 1.

The β+decay is possible when the mass (energy) constraint is fulfilled:

or for atomic masses:

3. Electron capture (Electron capture, known also as inverse beta decay is sometimes included as a type of beta decay, because
the basic nuclear process, mediated by the weak interaction, is the same.)

, mass number A do not change, atomic number Z decrease by 1.

Electron capture is possible when the mass (energy) constraint is fulfilled:

or for atomic masses:

For some nuclei all three conditions are fulfilled at once, thus all 3 processes can take place simultaneously. For example for
isotope the probabilities are: β- (40%), β+ (20%), e-capture (40%)

If β+ decay is possible, thus e-capture also takes place, since it is more energetically favored. Opposite statement is not true: when
e-capture is allowed, the β+ decay can be forbidden. If both procesess take place, thus for light nuclei β+ is dominating, for heavy
nuclei the e-capture is prefered.
β-decay
Energy Spectrum of Beta Decay
In both alpha and gamma decay, the resulting particle (alpha particle or photon) has a narrow energy
distribution, since the particle carries the energy from the difference between the initial and final
nuclear states. For example, in case of alpha decay, when a parent nucleus breaks down spontaneously
to yield a daughter nucleus and an alpha particle, the sum of the mass of the two products does not
quite equal the mass of the original nucleus (Mass Defect). As a result of the law of conservation of
energy, this difference appears in the form of the kinetic energy of the alpha particle. Since the same
particles appear as products at every breakdown of a particular parent nucleus, the mass-difference
should always be the same, and the kinetic energy of the alpha particles should also always be the
same. In other words, the beam of alpha particles should be monoenergetic.

It was expected that the same considerations would hold for a parent nucleus breaking down to a daughter nucleus and  a beta particle. Because
only the electron and the recoiling daughter nucleus were observed beta decay, the process was initially assumed to be a two body process,
very much like alpha decay. It would seem reasonable to suppose that the beta particles would form also a monoenergetic beam.
But the reality was different. The spectrum of beta particles measured by Lise Meitner and Otto Hahn in 1911 and by Jean Danysz in 1913
showed multiple lines on a diffuse background, however. Moreover virtually all of the emitted beta particles have energies below that predicted
by energy conservation in two-body decays. The electrons emitted in beta decay have a continuous rather than a discrete spectrum
appeared to contradict conservation of energy, under the assumption that beta decay is the simple emission of an electron from a nucleus. When
this was first observed, it appeared to be problematic for the survival of one of the most important conservation laws in physics!

To account for this energy release, Pauli proposed (in 1931) that there was emitted in the decay process another particle, later named by Fermi
the neutrino. It was clear, this particle must be highly penetrating and that the conservation of electric charge requires the neutrino to be
electrically neutral. This would explain why it was so hard to detect this particle. The term neutrino comes from Italian meaning “little neutral
one” and neutrinos are denoted by the Greek letter ν (nu). In the process of beta decay the neutrino carries the missing energy and also in this
process the law of conservation of energy remains valid.
β-decay
The first process is known as negative β decay and involves the creation and emission of an ordinary electron. The second
process is positive β decay or positron decay, in which a positively charged anti-electron (positron) is emitted. In the third
process, an atomic electron that strays too close to the nucleus is swallowed, allowing the conversion of a proton to a neutron.
In all three processes, yet another particle called a neutrino (or anti-neutrino ) is also emitted, but since the neutrino has no
electric charge, its inclusion in the decay process does not affect the identity of the other final particles.

In contrast to alpha decay, neither the beta particle nor its associated neutrino exist within the nucleus prior to beta decay, but are
created in the decay process. By this process, unstable atoms obtain a more stable ratio of protons to neutrons. The probability of
a nuclide decaying due to beta and other forms of decay is determined by its nuclear binding energy. For either electron or
positron emission to be energetically possible, the energy release must be positive.

According to Standard Model, Beta decay is governed by the weak interaction. During beta decay one of two
down quarks changes into an up quark by emitting a W– boson (carries away a negative charge). The W – boson then decays into
a beta particle and an antineutrino. This process is equivalent to the process, in which a neutrino interacts with a neutron.
β-decay
Conservation Laws in Beta Decay
In analyzing nuclear reactions, we apply the many conservation laws. Nuclear reactions are subject to classical conservation
laws for charge, momentum, angular momentum, and energy (including rest energies).  Additional conservation laws, not
anticipated by classical physics, are:
Law of Conservation of Lepton Number
Law of Conservation of Baryon Number
Law of Conservation of Electric Charge
Conservation Laws
In analyzing nuclear decay reactions, we apply the many conservation laws. Nuclear decay reactions are subject to
classical conservation laws for charge, momentum, angular momentum, and energy (including rest energies).
Conservation of energy. Energy, including rest mass energy, is conserved in nuclear reactions. Thus if parent nucleus is at
rest, we have: , where is the mass of initial (parent) nucleus, is the mass of final (daughter) nucleus, are the masses of
other particles born during decay, is the kinetic energy of all particles.

The spontaneous decay is possible just if .

This is indispensible condition, but also other conservation laws should be satisfied.
 
Additional conservation laws, not anticipated by classical physics, are:
•Law of Conservation of Lepton Number
•Law of Conservation of Baryon Number
•Law of Conservation of Electric Charge
Law of Conservation of Electric Charge
Law of Conservation of Electric Charge

In physics, there are two very important principles concerning the electric charge.

First is the law of conservation of electric charge. This law states that:

The algebraic sum of all the electric charges in any closed system is constant.
The only way to change the net charge of a system is to bring in charge from elsewhere, or remove charge from the
system. Charge can be created and destroyed, but only in positive-negative pairs.

Conservation of charge is thought to be a universal conservation law. No experimental evidence for any violation of this
principle has ever been observed. In particle physics, charge conservation means that in elementary particle reactions
that create charged particles, equal numbers of positive and negative particles are always created, keeping the net
amount of charge unchanged. Even in high-energy interactions in which particles are created and destroyed, such as the
creation of positron-electron pairs, the total charge of any closed system is exactly constant.

The second important principle is:


The magnitude of charge of the electron or proton is a natural unit of charge.
We say that charge is quantized. That is, every observable amount of electric charge is always an integer multiple of this
basic unit. This unit is called the elementary charge, e, approximately equal to 1.602×10−19 coulombs (except for particles
called quarks, which have charges that are integer multiples of  1⁄3e).
Law of Conservation of Baryon Number
What is Baryon Number? Baryons are composite particles made of three quarks, as opposed to mesons, which are composite
particles made of one quark and one antiquark.

Hyperon is any baryon containing


one or more strange quarks, but
no charm, bottom, or top quark.

In particle physics, a kaon, also called a K meson and denoted


K, is any of a group of four mesons distinguished by a quantum
number called strangeness. In the quark model they are
understood to be bound states of a strange quark (or antiquark)
and an up or down antiquark (or quark).
Law of Conservation of Baryon Number
In particle physics, the baryon number is used to denote which particles are baryons and which particles are not. Each baryon
has a baryon number of 1 and each antibaryon has a baryon number of -1. Other non-baryonic particles have a baryon number
of 0. The general definition of baryon number:

where  is the number of quarks, and  is the number of antiquarks. (So, quark baryon number =1/3).
The baryon number is a conserved quantum number in all particle reactions. The term conserved means that the sum of the
baryon number of all incoming particles is the same as the sum of the baryon numbers of all particles resulting from the
reaction.

Baryon number is a generalization of nucleon number, which is conserved in nonrelativistic nuclear reactions and decays.
The law of conservation of baryon number states that:
The sum of the baryon number of all incoming particles is the same as the sum of the baryon numbers of all particles resulting
from the reaction.
For example, the following reaction has never been observed:

even if the incoming proton has sufficient energy and charge, energy, and so on, are conserved. This reaction does not conserve
baryon number since the left side has B =+2, and the right has B =+1.
On the other hand, the following reaction (proton-antiproton pair production) does conserve B and does occur if the incoming
proton has sufficient energy (the threshold energy = 5.6 GeV):
β-decay
Conservation Laws in Beta Decay
In analyzing nuclear reactions, we apply the many conservation laws. Nuclear reactions are subject to classical conservation
laws for charge, momentum, angular momentum, and energy (including rest energies).  Additional conservation laws, not
anticipated by classical physics, are:
Law of Conservation of Lepton Number
Law of Conservation of Baryon Number
Law of Conservation of Electric Charge
Law of Conservation of Lepton Number
What are Leptons
A lepton is an elementary, half-integer spin (spin  1⁄2) particle that does not undergo strong interactions. Particles that do
participate in strong interactions are called hadrons (baryons and mesons , composed by quarks).

There are six leptons in the present structure, the electron, muon, and tau particles and their associated neutrinos.
Leptons are said to be elementary particles; that is, they do not appear to be made up of smaller units of matter. They
behave as point-like particles. All leptons are fermions, i.e. leptons are spin- 1⁄2 particles and thus that they are subject to
the Pauli exclusion principle. This fact has key implications for the building up of the periodic table of elements.

In particle physics, the lepton number is used to denote which particles are leptons and which particles are not.
Each lepton has a lepton number of 1 and each antilepton has a lepton number of -1. Other non-leptonic particles have a
lepton number of 0. The lepton number is a conserved quantum number in all particle reactions. A slight asymmetry in the
laws of physics allowed leptons to be created in the Big Bang.
Law of Conservation of Lepton Number
Two main classes of leptons exist:

Charged leptons. Charged leptons can combine with other particles to form


various composite particles such as atoms and positronium.

Electron. The electron is a negatively charged particle with a mass that


is approximately 1/1836 that of the proton. Electrons are located in an
electron cloud, which is the area surrounding the nucleus of the atom.
The electron is only one member of a class of elementary particles,
which forms an atom.

Muon. The muon is an elementary particle similar to the electron, with


an electric charge of −1 e and a spin of ½. Muons are heavier, having
more than 200 times as much mass as electrons. The muon is an
unstable subatomic particle with a mean lifetime of 2.2 µs.

Tau. The tau (τ), also called the tau lepton, tau particle, or tauon, is an
elementary particle similar to the electron, with an electric charge of
−1 e and a spin of ½. Taus are approximately 3,700 times more massive
than electrons. Tau leptons have a lifetime of 2.9×10−13 s.
Law of Conservation of Lepton Number
Two main classes of leptons exist:

Neutral leptons (neutrinos) are electrically neutral particles that rarely


interact with anything, and are consequently rarely observed. A
neutrino is an elementary subatomic particle with infinitesimal
mass (less than 0.3 eV..?) and with no electric charge. Neutrinos
are weakly interacting subatomic particles with ½ unit of spin.

Electron neutrino. The electron neutrino is a subatomic lepton


elementary particle which has the symbol νe. It has no net electric
charge and a spin of ½. Together with the electron it forms the
first generation of leptons, hence the name electron neutrino.

Muon neutrino. The muon neutrino is a subatomic lepton


elementary particle which has the symbol  νμ. It has no net
electric charge and a spin of ½. Together with the muon it forms
the second generation of leptons, hence the name muon
neutrino.

Tau neutrino. The tau neutrino is a subatomic lepton elementary


particle which has the symbol  ντ. It has no net electric charge and
a spin of ½. Together with the tauon it forms the third generation
Law of Conservation of Lepton Number
Conservation of Lepton Number – Electron Capture
Consider the electron capture mode. The reaction involves only first generation leptons, i.e.
electrons and neutrinos:

The antineutrino cannot be emitted, because in this case the conservation law would not be


fulfilled. The particle emitted with the neutron must be a neutrino.

Conservation of Lepton Number – Neutron Decay


Consider the decay of the neutron. The reaction involves only first generation leptons:
electrons and neutrinos:

Since the lepton number must be equal to zero on both sides and it was found that the
reaction is a three-particle decay ,  the third particle must be an electron antineutrino.

Conservation of Lepton Number – Muon Decay


The observation of the following decay reaction leads to the conclusion that there is a
separate lepton number for muons which must also be conserved.
Passage of Light charged particles (electrons) through matter
Summary of types of interactions:
1. Collisions with atomic electrons (Excitation and Ionization)
2. Bremsstrahlung (Elastic scattering off nuclei).
3. Cherenkov radiation.
4. Annihilation (only positrons)

Electrons (also positrons) interact through Coulomb scattering from atomic electrons, just like heavy charged particles. There
are, however, a number of important differences:

(1) Electrons, particularly those emitted in β decay, travel at relativistic speeds.

(2) Electrons will suffer large deflections in collisions with other electrons, and therefore will follow erratic paths. The range
(defined as the linear distance of penetration into the material) will therefore be very different from the length of the path that
the electron follows.

(3) In head-on collisions of one electron with another, a large fraction of the initial energy may be transferred to the struck
electron. (In fact, in electron-electron collisions we must take into account the identity of the two particles; after the collision,
we cannot tell which electron was incident and which was struck.)

(4) Because the electron may suffer rapid changes in the direction and the magnitude of its velocity, it is subject to large
accelerations, and accelerated charged particles must radiate electromagnetic energy. Such radiation is called bremsstrahlung
(German for "braking radiation").
Passage of Light charged particles (electrons) through matter
The bremsstrahlung  is electromagnetic radiation produced by the acceleration or deceleration
of a charged particle when deflected by magnetic fields (an electron by magnetic field of
particle accelerator) or another charged particle (an electron by an atomic nucleus). The name
bremsstrahlung comes from the German. The literal translation is ‘braking radiation’. From
classical theory, when a charged particle is accelerated or decelerated, it must radiate energy.
The bremsstrahlung is one of possible interactions of light charged particles with matter
(especially with high atomic numbers).
The two commonest occurrences of bremsstrahlung are possible:
Deceleration of charged particle. When charged particles enter a material they are decelerated
by the electric field of the atomic nuclei and atomic electrons.
Acceleration of charged particle. When ultra-relativistic charged particles move
through magnetic fields they are forced to move along a curved path. Since their direction of
motion is continually changing, they are also accelerating and so emit bremsstrahlung, in this
case it is referred to as synchrotron radiation.

Since the bremsstrahlung is much stronger for lighter particles, this effect is much more
important for beta particles than for protons, alpha particles, and heavy charged nuclei (
fission fragments). This effect can be neglected at particle energies below about 1 MeV,
because the energy loss due to bremsstrahlung is very small. Radiation loss starts to become
important only at particle energies well above the minimum ionization energy. At relativistic
energies the ratio of loss rate by bremsstrahlung to loss rate by ionization is approximately
proportional to the product of the particle’s kinetic energy and the atomic number of the
absorber.
Bremsstrahlung dependence on particle mass (add information)
Since the bremsstrahlung is much stronger for lighter particles, this effect is much more important for beta particles than for
protons, alpha particles, and heavy charged nuclei (fission fragments).

The total radiated power is[2] where (the velocity of the particle divided by the speed of light), is the Lorentz factor,

signifies a time derivative of , and q is the charge of the particle. In the case where velocity is parallel to
acceleration (i.e., linear motion), the expression reduces to[3]

where is the acceleration. For the case of acceleration perpendicular to the velocity ( ),
for example in synchrotrons, the total power is

Power radiated in the two limiting cases is proportional to or . Since , we see

that for particles with the same energy the total radiated power goes as or , which accounts for why electrons lose energy to
bremsstrahlung radiation much more rapidly than heavier charged particles (e.g., muons, protons, alpha particles). This is the reason a TeV
energy electron-positron collider (such as the proposed International Linear Collider) cannot use a circular tunnel (requiring constant

acceleration), while a proton-proton collider (such as the Large Hadron Collider) can utilize a circular tunnel. The electrons lose energy due to
bremsstrahlung at a rate times higher than protons do.
Passage of Light charged particles (electrons) through matter
As it was already told, beta particles can interact also via electron-nuclear interaction (elastic scattering off nuclei), which
can significantly change the direction of beta particle. Therefore their path is not so straightforward. The beta particles
follow a very zig-zag path through absorbing material, this resulting path of particle is longer than the linear penetration
(range) into the material.

Beta particles also differ from other heavy charged particles in the fraction of energy lost by radiative process known as
the bremsstrahlung. From classical theory, when a charged particle is accelerated or decelerated, it must radiate energy and
the deceleration radiation is known as the bremsstrahlung (“braking radiation”).

There is another mechanism by which beta particles loss energy via production of electromagnetic radiation. When the beta
particle moves faster than the speed of light (phase velocity) in the material it generates a shock wave of electromagnetic
radiation known as the Cherenkov radiation.

Positrons interact similarly with matter when they are energetic. But when the positron comes to rest, it interacts with a
negatively charged electron, resulting in the annihilation of the electron-positron pair.
Passage of Light charged particles (electrons) through matter
The expressions for the energy loss per unit path length for electrons were also derived by Bethe, and can be written in a
form similar as for heavy particles
For the “collisional losses” (due to excitation and ionization ) we have:

And for radiative losses (Bremsstrahlung):

Where T is the particle kinetic energy, β=v/c, 𝑍 is the medium atomic number, 𝜌 is the medium density, 𝑁𝐴 is Avogadro
number, is the atomic weight (moll mass) of the medium. )

The total energy loss is just the sum of these two contributions:
Passage of Light charged particles (electrons) through matter
An electron loses energy by bremsstrahlung at a rate nearly proportional to its energy T, while the ionization loss rate varies only
logarithmically with the electron energy T. The critical energy Ec is defined as the energy at which the two loss rates are equal.
The value of the critical energy when bremsstrahlung starts to prevail over ionization mechanism can be obtained by the
expressions: for solids and liquids and for gases.
Bremsstrahlung radiation is emitted when the fast-moving charged particle is decelerated in the Coulomb field of the atoms.
Though radiation takes place mainly due to the field of the nuclei, atomic electrons also contribute to the process. Since the
probability of the bremsstrahlung process is proportional to 1⁄m 2 (m being the mass of the particle), bremsstrahlung becomes the
dominant process in the interaction of the lightest charged particles (i.e. electrons and positrons) with most materials already
above a few tens of MeV. The bremsstrahlung process probability increases with Z2.
For electrons and positrons, which are much lighter than muons or protons, radiative losses are already important at much lower
energies. This is illustrated in Fig. , which shows electronic (ionization) and radiative mass stopping powers of electrons in
carbon and lead. Radiative losses start to dominate above a few tens of MeV in low-Z materials, and above a few MeV in high-Z
materials.

Fig. Mass stopping powers of electrons


in (left) carbon and (right) lead.
Passage of Light charged particles (electrons) through matter:
Cherenkov r.
Charged particles undergo another form of energy loss in media if they travel faster than the local speed of light. They emit the
Cherenkov radiation. It is electromagnetic radiation emitted when a charged particle (such as an electron) moves through a
dielectric medium faster than the phase velocity of light in that medium. It is similar to the bow wave produced by a boat
travelling faster than the speed of water waves. Cherenkov radiation occurs only if the particle’s speed is higher than the phase
velocity of light in the material. Even at high energies the energy lost by Cherenkov radiation is much less than that by the
other mechanisms (collisions, bremsstrahlung).
Cherenkov radiation can be used to detect high-energy charged particles (especially beta particles). In nuclear reactors or in a
spent nuclear fuel pool, beta particles (high-energy electrons) are released as the fission fragments decay. The glow is visible
also after the chain reaction stops (in the reactor). The Cherenkov radiation can characterize the remaining radioactivity of spent
nuclear fuel, therefore it can be used for measuring of fuel burnup.

1
=¿
𝑛
Passage of Light charged particles (electrons) through matter:
annihilation
The coulomb forces that constitute the major mechanism of energy loss for electrons are present for either positive or
negative charge on the particle and constitute the major mechanism of energy loss also for positrons. Whatever the interaction
involves a repulsive or attractive force between the incident particle and orbital electron (or atomic nucleus), the momentum
and energy transfer for particles of equal mass are about the same. Therefore positrons interact similarly with matter when
they are energetic. The track of positrons in material is similar to the track of electrons. Even their specific energy loss and
range are about the same for equal initial energies.

At the end of their path, positrons differ significantly from electrons. When a positron (antimatter particle) comes to rest, it
interacts with an electron (matter particle), resulting in the annihilation of the both particles and the complete conversion of
their rest mass to pure energy (according to the E=mc2 formula) in the form of two oppositely directed 0.511 MeV gamma
rays (photons).
Electron–positron annihilation occurs when a negatively charged electron and a
positively charged positron collide. When a low-energy electron annihilates a low-
energy positron (antiparticle of electron), they can only produce two or more photons
(gamma rays). The production of only one photon is forbidden because of
conservation of linear momentum and total energy. The production of another particle
is also forbidden because of both particles (electron-positron) together do not carry
enough mass-energy to produce heavier particles. When an electron and a positron
collide, they annihilate resulting in the complete conversion of their rest mass to pure
energy (according to the E=mc2 formula) in the form of two oppositely directed 0.511
MeV gamma rays (photons).
e− + e+ → γ + γ (2x 0.511 MeV)
γ-decay
Gamma decay or γ decay represents the disintegration of a parent nucleus to a daughter through the emission of gamma
rays (high energy photons). This transition (γ decay) can be characterized as:

As can be seen, if a nucleus emits a gamma ray, atomic and mass numbers of daughter nucleus remain the same, but daughter
nucleus will form different energy state of the same element. Note that, nuclides with equal proton number and equal mass
number (thus making them by definition the same isotope), but in a different energy state are known as nuclear isomers. We
usually indicate isomers with a superscript m, thus: 241mAm or 110mAg.

In most practical laboratory sources, the excited nuclear states are created in
the decay of a parent radionuclide, therefore a gamma decay
typically accompanies other forms of decay, such as alpha or beta decay.
Typically after a beta decay (isobaric transition), nuclei usually contain too
much energy to be in its final stable or daughter state.

Gamma rays are high-energy photons with very short wavelengths and thus


very high frequency. Gamma rays from radioactive decay are in the energy
range from a few keV to ~8 MeV, corresponding to the typical energy levels in
nuclei with reasonably long lifetimes.
γ-decay
Another example of gamma ray production due to radionuclide decay is the decay scheme for cobalt-60, as illustrated in the
accompanying diagram. First, 60Co decays to excited 60Ni by beta decay emission of an electron of 0.31 MeV. Then the
excited 60Ni decays to the ground state (see nuclear shell model) by emitting gamma rays in succession of 1.17 MeV
followed by 1.33 MeV. This path is followed 99.88% of the time:
γ-decay
In contrast to alpha and beta radioactivity, gamma radioactivity is governed by an electromagnetic interaction rather than
a weak or strong interaction. As in atomic transitions, the photon carries away at least one unit of angular momentum (the
photon, being described by the vector electromagnetic field, has spin angular momentum of ħ.Gamma decay may
follow nuclear reactions such as neutron capture, nuclear fusion, or nuclear fission. Most of nuclear reactions produce
extremely unstable nuclei that decay as soon as they are formed in nuclear reactions (half-life less than 10 -11s) and are not
generally classified as nuclear isomers. Moreover, these nuclei usually produce a cascade of gamma rays and the gamma-ray
cascade concludes when all excess energy of the excited nucleus is released.

In certain cases, the excited nuclear state that follows the emission of a beta particle or other type of excitation, are able
to stay in metastable state for a long time (hours, days and sometimes much longer) before undergoing gamma decay, in
which they emit a gamma ray. These long-lived excited nuclei are known as isomeric states (or isomers) and their decays are
termed isomeric transitions. The process of isomeric transition is therefore similar to any gamma emission, but differs in
that it involves the intermediate metastable excited state(s) of the nuclei.

Metastable nuclei are often characterized by high nuclear spin, requiring a change in spin of several units or more with
gamma decay, instead of a single unit transition that occurs in only 10 −12 seconds. The rate of gamma decay is also slowed
when the energy of excitation of the nucleus is small. An example is the decay of the isomer or metastable state of
protactinium:
Extremely unstable nuclei that decay as soon as they are formed in nuclear
reactions (half-life less than 10-11s) are not generally classified as nuclear
isomers. Isomeric transitions must occur by higher order multipole
transitions (in contrast to gamma emission that occurs by dipole radiation)
that occur on a longer time-scale. E=0.0698Mev
γ-decay
Internal Conversion

Internal conversion is an electromagnetic process, by which


a nuclear excited state decays by the direct emission of one
of its atomic electrons. Internal conversion competes
with gamma emission, but in this case the electromagnetic
multipole fields of the nucleus do not result in the emission of
a gamma ray, instead, the fields interact directly with atomic
electrons.

In contrast to beta decay, which is governed by a weak force,


the electron is emitted from the radioactive atom, but not from
the nucleus. For this reason, internal conversion is possible
whenever gamma decay is possible, except in the case where
the atom is fully ionised.
γ-decay energy of the excited nucleus is released (virtual photon).

Internal Conversion

As can be seen, if a nucleus decays via internal


conversion, atomic and mass numbers of daughter nucleus remain
the same, but daughter nucleus will form different energy state of
the same element. This is very similar to gamma decay, but in this
case, there is no gamma ray emitted from an excited nucleus.

Since the process leaves a vacancy in the electron energy level from
which the electron came, the outer electrons of the atom cascade
down to fill the lower atomic levels, and one or more characteristic
X-rays are usually emitted. Sometimes X-ray may interact with
another orbital electron, which may be ejected from the atom. This
second ejected electron is called an Auger electron (see Atom
Physics section, X-ray spectra).

This is very similar to electron capture, but in case of electron


capture, a nucleus changes its atomic number. As a result, the atom
thus emits primary high energy electron, characteristic X-rays or
secondary Auger electron, none of which originate in that nucleus.
γ-decay
Internal Conversion
In the quantum mechanical model of the electron, there is a finite
probability of finding the electron within the nucleus. During the
internal conversion process, the wavefunction of the K shell electron
(inner shell electron) is said to penetrate the volume of the atomic
nucleus. Note that, typical nuclear radii are of the order 10-14 m. In
this case, the electron may couple to an excited nucleus and take the
energy of the nuclear transition directly, without an intermediate
gamma ray.

Therefore, most internal conversion electrons (ICE) come from the


K shell, as these electrons have the highest probability of being
within the nucleus. However, the s states in the L, M, and N shells
are also able to couple to the nuclear fields and cause ICE ejections
from those shells.

The internal conversion electron (ICE) energy, is the transition


energy, Etransition, minus the binding energy of the orbital electron, E b.e.,
as:
γ-decay
For example, 203Hg (Mercury) is beta radioactive nuclide, which produces a continuous beta spectrum with maximum energy 214 keV. This
decay produces an excited state of the daughter nucleus 203Tl, which then decays very quickly (~ 10 -10 s) to its ground state emitting a gamma
ray of energy 279.2 keV or an internal conversion electron. If we analyze a spectrum of beta particles, we can see the typical continuous
spectrum of beta particles as well as narrow peaks at specific energies. These peaks are produced by internal conversion electrons (ICE).
Since the binding energy of the K electrons in 203Tl (Thallium) amounts to 85.5 keV, the K line has an energy of: T e (K) = 279.2 – 85.5 = 194
keV.

Because of lesser binding energies, the L- and M-lines have higher energies. Since the internal conversion process can interact with any of the
orbital electrons, the result is a spectrum of internal conversion electrons which will be seen as superimposed upon the electron energy spectrum
of the beta emission. These relative intensities of these ICE peaks can give information about the electric multipole character of the nucleus and
about the decay process.
Gamma Ray Attenuation
The intensity of the gamma and X rays is decreasing during the propagation in medium
proportionally to medium particle density (atom density) n and medium thickness dx: .
By introducing coefficient of proportionality we can write: , where is the beam total
effective attenuation cross-section. After integration we have:

The product is called linear attenuation (or absorption) coefficient.

Gamma rays and X rays interact with matter primarily through three processes:

Photoelectric effect,
Compton scattering,
Electron-positron pair production.

In the photo­electric effect, a photon is absorbed by an atom and one of the atomic
electrons, known in this case as a photoelectron, is released. (Free electrons cannot
absorb a photon and recoil. Energy and momentum cannot both be conserved in such a
process; a heavy atom is necessary to absorb the momentum at little cost in energy.) The
kinetic energy of the electron is equal to the photon energy less the binding energy of
the electron:
Gamma Ray Attenuation
The probability for photoelectric absorption is difficult to calculate, but from
experimental studies we know several features:

it is most significant for low­energy photons ( ~ 100 keV),

it increases rapidly with the atomic number Z of the absorber atoms (roughly as n=4÷5)
and it decreases rapidly with increasing photon energy (roughly as )

Furthermore, there are in the probability (cross-section ) for photoelectric absorption


discontinuous jumps at energies corresponding to the binding energies of particular
electronic shells. That is, the binding energy of a K-shell electron in Pb is 88 keV.
Incident photons of energy less than 88 keV cannot release K-shell photoelectrons
(although they can release less tightly bound electrons from higher shells). When the
photon energy is increased above 88 keV, the availability of the K-shell electrons to
participate in the photoelectric absorption process causes a sudden increase in the
absorption probability, known as the K-absorption edge or simply K edge. Figure shows a
sample of the photoelectric absorption cross section.
Figure. Photoelectric cross section per single atom in Pb.
The discrete jumps correspond to the binding energies of
various electron shells; the K-electron binding energy, for
example, is 88 keV. To convert the cross section to the
linear absorption coefficient in cm-1, multiply by 0.033.
Gamma Ray Attenuation
Compton scattering is the process by which a photon scatters from a nearly free atomic
electron, resulting in a less energetic photon and a scattered electron carrying the
energy lost by the photon. Figure shows a schematic view of the process. If we regard
the struck electron as free and at rest (a good approxima­tion, since the photon energy is
usually large compared with the orbital energies of the loosely bound outer atomic
electrons), then conservation of linear momentum and total energy (using relativistic
dynamics) gives:

If we observe the scattered photon, then we may eliminate the unobserved variables and , giving the Compton-
scattering formula

The probability for Compton scattering at an angle θ can be determined through a quantum mechanical calculation of the
process. The result is the Klein-Nishina formula for the differential cross section per electron:
Gamma Ray Attenuation
If we are interested in the absorption of photons (that is, their removal from the incident beam of photons), we must integrate
over all angles since we do not observe the scattered photon. The result is for each electron in the scattering medium:

is the photon energy in units of the electron rest energy, is the classical electron radius

The third interaction process is pair production, in which a photon


creates an electron-positron pair; the photon disappears in this process.
The energy balance is

where and are the energies of the positron and electron. Like
photoelectric absorption, this process requires the nearby presence of a
massive atom for momentum conservation, but the recoil energy given
to the atom is negligible compared with the other terms. There is
obviously a threshold of , or 1.022 MeV, for this process, and in
general pair production is important only for photons of high energy.
The pair production becomes dominant only for energies above 5
MeV.
Gamma Ray Attenuation
The total probability per unit length , for removal of a photon is called the total linear attenuation coefficient; it is simply the
sum of the respective probabilities for photoelectric absorption:

(since nZ is the electron density)

Figure shows some representative values for the energy


dependence of the attenuation coefficients.

Photon mass attenuation coefficients (), equal to the linear


attenuation coefficients divided by the density (to suppress
effects due simply to the number of electrons in the
material) for the three processes in Al and Pb.
Gamma Ray Attenuation
The half value layer expresses the thickness of absorbing material needed for reduction of
the incident radiation intensity by a factor of two. There are two main features of the half
value layer:
•The half value layer decreases as the atomic number of the absorber increases. For
example 35 m of air is needed to reduce the intensity of a 100 keV gamma ray beam by a
factor of two whereas just 0.12 mm of lead can do the same thing.
•The half value layer for all materials increases with the energy of the gamma rays. For
example from 0.26 cm for iron at 100 keV to about 1.06 cm at 500 keV.

When characterizing an absorbing material, we can use sometimes the mass attenuation coefficient.  The mass attenuation
coefficient is defined as the ratio of the linear attenuation coefficient and absorber density (μ/ρ). The attenuation of gamma
radiation can be then described by the following equation: I=I0.e-(μ/ρ).ρl, where ρ is the material density, (μ/ρ) is the mass
attenuation coefficient and ρ.l is the mass thickness.

The measurement unit used for the mass attenuation coefficient cm 2g-1.For intermediate energies the Compton scattering
dominates and different absorbers have approximately equal mass attenuation coefficients. This is due to the fact that cross
section of Compton scattering is proportional to the Z (atomic number) and therefore the coefficient is proportional to the
material density ρ. At small values of gamma ray energy or at high values of gamma ray energy, where the coefficient is
proportional to higher powers of the atomic number Z (for photoelectric effect σ f ~ Z5; for pair production σp ~ Z2), the
attenuation coefficient μ is not a constant.
Accelerators
The first historical particle accelerator was built by the Wilhelm Conrad Röntgen (1900). It consisted in a vacuum tube
containing a cathode connected to the negative pole of a DC voltage generator. Electrons emitted by the heated cathode were
accelerated while flowing to another electrode connected to the positive generator pole (anode). Collisions between energetic
electrons and anode produced X-rays.

The development of charged-particle accelerators has progressed along double paths which by the appearance of particle
trajectories are distinguished as linear accelerators and circular accelerators. Particles in linear accelerators travel on a straight
line and pass only once through the accelerator structure while in a circular accelerator they follow a closed orbit periodically
for many revolutions accumulating energy at every passage of the accelerating structure.

No fundamental advantage or disadvantage can be claimed for one or the other class of accelerators. It is mostly the particular
application and sometimes the available technology that determines the choice between both classes. Both types have been
invented and developed throughout the twentieth century, and continue to be improved and optimized as associated technologies
advance.
Electrostatic Accelerators
In electrostatic accelerators the potential difference between two electrodes is
used for particle acceleration as shown in Fig. The most simple such
arrangement has been used now for almost two centuries in glow discharge
tubes for fundamental research on the nature of plasmas, as light sources or as
objects of aesthetic interest due to colorful phenomena in such tubes. In
another, more modern application electrons are accelerated in an x-ray tube by
high electrostatic fields and produce after striking a metal target intense x-rays
used in medicine and industry.

The voltages that can be achieved by straight voltage transformation and rectifi-
cation are quite limited by electrical breakdown effects to a few 10,000 V/cm.
More sophisticated methods of producing high voltages therefore have been
developed to reach potential differences of up to several million volts.

To distribute evenly the electric fields of high potential differences a series of


irises are distributed along the acceleration column and separated by
appropriate resistors to break down the high voltage into smaller steps between
the irises.

A variety of techniques to obtain high voltages have been developed and


applied to particle acceleration with more or less success.
Electrostatic Accelerators
Cascade Generators
 
The basic method implemented in the cascade generator is that of a voltage
multiplier circuit which has been proposed by Greinacher in 1914 and Schenkel
in 1919 which allows to achieve a multiplication of the voltage across the
plates of a capacitor.

A set of capacitors are charged through appropriately placed diodes from an


alternating current source (Fig. ) in such a way that during the positive half
wave, half the capacitors are charged to a positive voltage and during the
negative half wave, the other half of the capacitors are charged to a negative
voltage thus providing twice the maximum ac voltage. By arranging 2N
capacitors in this way the charging voltage can be multiplied by the factor 2N.

Based on this method Cockcroft and Walton developed appropriate high-


voltage techniques and built the first high energy particle accelerator reaching
voltages as high as several million Volt. Applying the high voltage to a beam of
protons they were able for the first time to initiate through artificially
accelerated protons a nuclear reaction. In this case it was the conversion of a
Lithium nucleus into two Helium nuclei.
750 kV Cockcroft-Walton
generator at CERN
Electrostatic Accelerators
1) HV generator based on a system of multiple rectifiers

2) At point A a transformer produces U(t) = U sin(ωt)

3) The first rectifying diode ensures at point B U never goes negative ⇒ C1


charges up to U

4) At point B the voltage oscillates between 0 and 2U. C2 is charged up to 2U.


The third diode ensures in potential in C does not fall below 2U and so on...

5) Maximum achievable voltage is 2NU, with N number of the rectifier stages

6) Voltages up to about 4 MV can be reached

7) Such Cockcroft-Walton accelerators turned out to be very efficient and are


still used as the first step in modern proton accelerator systems. Obviously
with this kind of voltage generation it is not possible anymore to produce a
continuous stream of particles. Because of the switching process, there is a
time to charge the capacitors followed by a time to apply the multiplied
voltage to particle acceleration. As a consequence, we observe a pulsed
particle beam from a Cockroft-Walton accelerator.
Electrostatic Accelerators
Van de Graaff Accelerator
 Much higher voltages can be reached with a Van de Graaff accelerator. Here,
electric charge is extracted by field emission from a pointed metal electrode
and sprayed onto an isolated endless belt. This belt is moved by motor action to
carry the charge to the inside of a hollow sphere (conducting dome), where the
charge is stripped off again by reverse field emission onto a pointed metal
electrode which is connected to the inside of the sphere.

The principle of this electrostatic generator is shown in Fig. Electrical charges


in a metallic conductor collect on the outside and it is therefore possible to
continuously accumulate electrical charge by deposition to the inside surface of
a hollow metallic sphere. If the whole system is placed into a high pressure
vessel filled with an electrically inert gas like Freon or SF6; voltages as high as
20 MV can be reached.

  High electrostatic voltages from a Van de Graaff generator cannot be applied


directly to just two electrodes. Because of the great distance between the
electrodes necessary to avoid voltage break down the fields would not be
distributed uniformly along the axis of the acceleration column.
Therefore, the voltage is applied to a series of resistors connected to iris
electrodes which allow a uniform distribution of the electrical field along the
acceleration column as shown in Fig. – thus spark discharge risk highly reduced.
Electrostatic Accelerators
Van de Graaff Accelerator
   
The high voltage can be used to accelerate
electrons as well as protons or ions. In the latter two
cases more than double the accelerating voltage can
be achieved in a Tandem Van de Graaff accelerator.

If a proton beam must be accelerated, the


accelerating process would start with negatively
charged hydrogen ions H- from a plasma discharge
tube which are then accelerated say from ground
potential to the full Van de Graaff voltage +V. At
that point the two electrons of the negative
hydrogen ion are stripped away by a thin foil or gas
curtain resulting by charge exchange in positively
charged protons which can be further accelerated
between the potential +V and ground potential to a
total kinetic energy Ekin = e2V.
The 5 MeV Van de Graaff generator built in 1937 by the
LIMITS OF ELECTROSTATIC ACCELERATORS: Westinghouse Electric company in Forest Hills, Pennsylvania
The main limit in the achievable voltage is the
breakdown due to insulation problems. www.youtube.com/watch?v=Q9bijrQfS6E
Linear Accelerators (linacs )
Acceleration by rf Fields
 The most successful acceleration of particles is based on the use of rf fields for which by now powerful sources exist.
The principle of the linear accelerator based on microwave fields and drift tubes was proposed by Ising and Wideroe.

The RF system converts the high voltage pulses from the modulator into pulsed radio frequency energy. The RF pulses are sent
to the accelerating structure to setup an electric field which is used for charged particle acceleration. The main component of a
RF system is the microwave source. There is a variety of microwave tubes for generating and amplifying microwave signals. The
two most common ones used in linacs are magnetrons and klystrons.
Linear Accelerators
Acceleration by rf Fields
 The principle of the linear accelerator based on microwave fields and drift tubes was proposed by Ising and Wideroe. The
accelerator consists of a series of coaxial metallic tubes where the accelerating field is generated in gaps between adjacent tubes.
In this method particles are accelerated by repeated application of rf fields. Wideroe constructed such an accelerator and was
able to accelerate potassium ions up to 50 keV.

While the principle is simple, the realization requires specific conditions to ensure that the particles are exposed to only
accelerating rf fields. The particles travel through the metallic tubes while the field is not suitable for acceleration as shown in
Fig. The tubes shield the particles from external rf fields and the length of the tube segments are chosen such that the particles
reach the gap between two successive tubes only when the rf field is accelerating.

Synchronicity Condition
For efficient acceleration the motion of the particles must be synchronized
with the rf fields in the accelerating sections. The distance between the center
of two adjacent gaps must be equal to the travel time of the particles from one
gap to the next. The length of the drift tubes are chosen such that the particles
travel for most of the rf period in the field free interior and emerge in a gap to
the next drift tube at a moment the field is accelerating. The length of the
shielding tubes is therefore almost as long as it take the particles to travel in a
full rf period . In this case, we have synchronism between particle motion and
rf field and the length of the i-th drift tube is where vi is the velocity of the
particles in the i-th tube and Trf the rf period.
Linear Accelerators
 Stimulated by the successful acceleration of potassium ions by Wideroe, a group led by Sloan and Lawrence at Berkeley were
able to build a 50 kV rf generator oscillating at 10 MHz and delivering a gap voltage of 42 kV. Applying this to 30 acceleration
tubes they were able to accelerate mercury ions to a total kinetic energy of 1.26 MeV.

In the 1920s when this principle was developed it was difficult to build high frequency generators at significant power. In 1928 rf
generators were available only up to about 7 MHz and numerical evaluation shows that this principle was useful only for rather
slow particles like low energy protons and ions. The drift tubes can become very long for low rf frequencies and particles
traveling with, for example, half the speed of light would require a drift tube length of 10.7 m at 7 MHz. Such long drift tubes add
up quickly to a very long accelerators before the particles approach the speed of light. To reduce the length of the tubes, higher
frequencies are required.

Further progress in the development of rf linear accelerators therefore depended greatly on the development of rf equipment at
high frequency which happened during World War II in connection with the development of radar systems. In 1937, Hansen and
the Varian brothers invented the klystron at Stanford. Soon the feasibility of high power klystrons had been established which to
this date is one of the most efficient rf amplifiers available. The first klystron was developed for 3,000 MHz which is still the
preferred frequency for high energy electron linear accelerators. The klystron principle is economically feasible from about 100
MHz to more than 10 GHz. With such a wide range of high frequencies available, the principle of rf acceleration in linear
accelerators has gained quick and continued prominence for the acceleration of protons as well as electrons.
Linear Accelerators
 Going to higher frequencies, however, the capacitive nature of the Wideroe structure becomes very lossy due to electromagnetic
radiation. To overcome this difficulty, Alvarez (1946) proposed to enclose the gaps between the tubes by metallic cavities (Fig.).
The acceleration section would now be composed of a series of tubes forming, together with the outer enclosure, a resonant
cavity. This Alvarez structure is still the preferred pre-accelerator to accelerate protons and ions from a few hundred keV out of a
Cockroft-Walton electrostatic generator to a few hundred MeV for injection into a booster synchrotron. Because of the lower
velocity of protons and ions at up to a few hundred MeV the operating frequency for proton linacs is generally around 200 MHz.

The electric field polarizes the drift tubes so that periodically the two ends
are charged to opposite sign.

When charging currents flow to the right along the tubes, an equal current
flows to the left along the inside wall of the outer cylinder.

The tubes are supported at their centers by radial rods along which no
current flows.

Alvarez DTL. Ions travel from gap to gap


in a full period of the rf field. Arrows
indicate the charging currents. www.youtube.com/watch?v=C79838wtRZo
Linear Accelerators

Alvarez DTL Illustrating the decreasing DT


diameter increasing with DT length to
maintain cavity resonance

Open Alvarez tank


Linear Accelerators

www.linearcollider.org
Circular Accelerators
Parallel with the development of electrostatic and linear rf accelerators the potential of circular accelerators was recognized and a
number of ideas for such accelerators have been developed over the years. Technical limitations for linear accelerators
encountered in the early 1920s to produce high-power rf waves

Interest in circular accelerators quickly moved up to the forefront of accelerator design and during the 1930s made it possible to
accelerate charged particles to many million electron volts. Only the invention of the rf klystron by the Varian brothers at
Stanford in 1937 gave the development of linear accelerators the necessary boost to reach par with circular accelerators again.

Circular accelerators are based on the use of magnetic fields to guide the charged particles along a closed orbit. The acceleration
in all circular accelerators but the betatron is effected in one or few accelerating cavities which are traversed by the particle
beam many times during their orbiting motion. This greatly simplifies the rf system compared to the large number of energy
sources and accelerating sections required in a linear accelerator. While this approach seemed at first like the perfect solution to
produce high energy particle beams, its progress soon became limited for the acceleration of electrons by copious production of
synchrotron radiation.

The simplicity of circular accelerators and the absence of significant synchrotron radiation for protons and heavier particles like
ions has made circular accelerators the most successful and affordable principle to reach the highest possible proton energies for
fundamental research in high energy physics. Protons are being accelerated into the TeV range in the Large Hadron Collider
(LHC) at CERN in Geneva, Switzerland.
For electrons the principle of circular accelerators has reached a technical and economic limit at about 28 GeV due to
synchrotron radiation losses, which make it increasingly harder to accelerate electrons to higher energies. Further progress in
the attempt to reach higher electron energies is being pursued through the principle of linear colliders, where synchrotron
radiation is avoided.
Betatron
The first “circular electron accelerator” has been invented and developed a
hundred years ago in the form of an electrical current transformer. Here we
find the electrons in the wire of a secondary coil accelerated by an electro
motive force generated by a time varying magnetic flux through the area
enclosed by the secondary coil. This idea was picked up independently by
several researchers.

The betatron makes use of the transformer principle, where the secondary
coil is replaced by an electron beam circulating in a closed doughnut shaped
vacuum chamber. A time-varying magnetic field is enclosed by the electron
orbit and the electrons gain an energy in each turn which is equal to the
electro-motive force generated by the varying magnetic field. The principle
arrangement of the basic components of a betatron are shown in Fig.

Different, more efficient accelerating methods have been developed for


protons, and betatrons are therefore used exclusively for the acceleration of
electrons as indicated by it’s name. Most betatrons are designed for modest
energies of up to 45 MeV and are used to produce electron and hard x-ray
beams for medical applications or in technical applications to, for example,
examine the integrity of full penetration welding seams in heavy steel
containers.
www.youtube.com/watch?v=W_i5WBgFbFo&t=5s
First betatron of Donald Kerst, 1940
Cyclotron
The next-generation accelerator after the linac is the cyclotron (Figure). The
beam is bent into a circular path by a magnetic field, and the particles orbit inside two
semicircular metal chambers called "dees" because of their shape. The dees are
connected to a source of alternating voltage. When the particles are inside the dees,
they feel no electric field and follow a circular path under the influence of the
magnetic field. In the gap between the dees, however, the particles feel an
accelerating voltage and gain a small energy each cycle.
The essential design idea of the cyclotron was conceived by Ernest Lawrence at the
University of California at Berkeley in 1929. The critical feature is that the time it takes
for a particle to travel one semicircular path is independent of the radius of the path-
as particles spiral to larger radii, they also gain energy and move at greater speed, and
the gain in path length is exactly compensated by the increased speed. If the half-
period of the ac voltage on the dees is set equal to the semicircular orbital time, then
the field alternates in exact synchronization with the passage of particles through the
gap, and the particle sees an accelerating voltage each time it crosses the gap.
The Lorentz force in the circular orbit, qvB, provides the necessary centripetal acceleration to maintain the circular motion at an
instantaneous radius r, and thus

and the time necessary for a semicircular orbit is , thus the frequency of the ac voltage is , which is often called the cyclotron
frequency or cyclotron resonance frequency for a particle of charge q and mass m moving in a uniform field B.
Cyclotron
Last equation shows that v and B are intimately linked-for a
given field strength, the frequency can only have a certain
value for resonance. The velocity increases gradually as the
particle spirals outward, and the greatest velocity occurs at the
largest radius R, according to above equation: and maximum
kinetic energy

The first practical cyclotron for accelerating particles was built


by Lawrence and M. Stanley Livingston at Berkeley in 1931.
The dees had a 12.5-cm radius and the cyclotron was able to
produce protons of energy 1.2 MeV in. a field of about 1.3 T
(13 kG); the corresponding frequency is about 20 MHz. Within
a few years, the radius had been extended to about 35 cm and
the basic particle energy to 10 MeV protons, 5 MeV deuterons,
and 10 MeV a's. By the end of the 1930s, radii of 75 cm had
been achieved, extending the range to 40 MeV a's and
protons, and 20 MeV deuterons. Figure 15.12 shows the 75-
cm cyclotron of the Argonne National Laboratory.
Cyclotron
As the beam in a cyclotron travels outward toward the edge of the machine, the magnetic
field lines are diverted somewhat from the true vertical (Figure). There are two effects of this
fringing field, one beneficial and one harmful. The curvature of the field lines gives a net
force component toward the median plane, which tends to provide focusing and to
counteract the tendency of the beam to diverge. At the same time, however, the field loses
its uniformity and the resonance condition (Equation ) can no longer be maintained if the
frequency is held constant.

A more serious difficulty comes from the relativistic behavior of the accelerated particles. If we account the relativistic mass

We have the ac frequency dependence on velocity . To maintain the resonance condition as v increases we must also increase
B, and so the field should be larger at the larger radii. This can be accomplished by "shimming" the field, which would then
show magnetic lines curving inward (opposite to those of Figure above) and would have a corresponding and undesirable
defocusing effect. In the basic design of the fixed-field, fixed-frequency cyclotron, there is no acceptable way to compensate
for the relativistic effect, and this provides an ultimate limit on the size of such machines. For protons, an energy of about 40
MeV is the maximum that can be achieved, corresponding to
Synchrocyclotron
To overcome this problem, one solution is to vary the frequency, resulting in a frequency-modulated cyclotron, called a
synchrocyclotron. It is obvious that in a variable-frequency cyclotron, a continuous beam is not possible, for the time to travel
the semicircular orbits will no longer be constant and equal to the half-period (which is now variable). Thus the particles travel
through the cyclotron in bunches, and the frequency is swept from its maximum value (when the bunch is near the center, the
particles are only slightly accelerated, and the relativistic increase in mass is slight) to its minimum value (when the bunch is
ready to exit the cyclotron, the maximum energy is attained, and the mass has its largest value). Particles in the bunches will
arrive at attained, and the mass has its largest value). Particles in the bunches will arrive at the gap between the dees at
different times. Phase stability provides a sort of time-focusing effect; those particles that arrive early are delayed somewhat
and on the next cycle are closer to the center of the bunch, while those that arrive late are advanced and likewise pushed
closer to the center.

The Berkeley synchrocyclotron, first operated in 1946, is the highest energy synchrocyclotron. The energy of extracted
protons is 740 MeV, and the field strength is about 2.3 T, corresponding to a frequency of about 20 MHz at the largest orbital
radius (where the mass is 1.8 times the rest mass). Other comparable synchrocyclotrons have operated at Dubna in the USSR
and at the European Center for Nuclear Research (CERN) in Geneva.

For more details

www.youtube.com/watch?v=L5zhpLfnqGc&list=RDCMUCOfLm6 www.youtube.com/watch?v=rOXfm6EezeA
gZGt3vwTMKRg-irhg&start_radio=1&rv=L5zhpLfnqGc
SYNCHROTRON
Extending the cyclotron or synchrocyclotron to higher energy
means building machines of larger radii. Because the magnet is the
principal factor in the cost of a cyclotron, we expect that costs of
building larger cyclotrons will scale roughly as the cube of the
energy. Extending the present generation of ~ 500 MeV cyclotrons
(costing of order $108) to the high-energy regime (even as low as 5
GeV, quite insufficient for the study of fundamental phenomena
of recent or current interest) quickly puts the cost into the realm
of the U.S. gross national product! The solution to this dilemma is
the synchrotron accelerator, in which both the magnetic field
strength and the resonant frequency are varied.

Figure shows the simplest design for a synchrotron. The essential feature that keeps the costs reasonable as the energy is
extended is that particles orbit at a very nearly constant radius at high energies. The magnetic field therefore need be applied
only at the circumference, not throughout the entire circular volume, as in an ordinary cyclotron. An annular structure magnet
accomplishes the task. Particles follow a circular path and are accelerated by a resonant electric field as they cross a gap during
each orbit. As the energy increases, the frequency of the ac voltage across the gap must increase to maintain the resonance;
simultaneously, the magnetic field must increase to keep the radius constant). In a magnetic field of strength B, a particle of
charge e moves in a circular arc of radius r at momentum p = qrB. The total relativistic energy of the particle is For a given r,
obtain the relationship between B and v necessary to maintain the synchronization.
SYNCHROTRON
The largest synchrotron-type accelerator,
also the largest particle accelerator in the
world, is the 27-kilometre-circumference
Large Hadron Collider (LHC) near Geneva,
Switzerland, built in 2008 by the European
Organization for Nuclear Research (CERN).
History
Usefull Energy: (in the Centre-of-Mass Frame = CM frame)
Consider the collision of two particles with energy, momentum
and rest mass (E , p , m ) , (E , p , m ) at an incident angle Ɵ.
1 1 1 2 2 2

The following quantity is an invariant (Lorentz invariant), since s=c2P2*P1 ,where Pi=(Ei/c,pi), Pi=(Ei/c,-pi) is 4-momentum of i-
particle:
s = (E1 +E2)2 - (c·p1+c·p2)2

being   p1  and  p2 particle momentum vectors  and   c = speed of light. Operating on the above equation we have:

s = (E12+ E22 + 2E1·E2) - (c2p12+c2p22 + 2p1c·p2c·cosƟ)


CM frame is defined as
a frame where total
momentum is zero 0= p1+p2 Rearranging the terms:

s = (E12 - c2p12) + (E22 - c2p22) + (2E1·E2 - 2p1c·p2c·cosƟ)

Since  (Ei2 - c2pi2) = (mi·c2)2   we have:


In CM frame we have s= (E1(cm) +E2(cm))2,
or in other words , s = (m1·c2)2 + (m2 c2)2 + 2(E1·E2 - p1c·p2c·cosƟ)   [1]
is the TOTAL ENERGY in CM frame
Usefull Energy: (in the Centre-of-Mass Frame = CM frame)
1.- Fixed target.

Let m1 be the projectile and  m2 the target.

We have: p2 = 0 and E2 = m2 c2 , in CM frame. Then equation [1] gives:

s = (m1·c2)2 + (m2·c2)2 + 2·E1· m2 c2

Take into account that E1 >> m1·c2 , m2 c2 , we have: s ~ 2E1·m2·c2

2.- Head-on collisions.
Mass m1 and m2 in a head-on collision (Ɵ = 180º)  cosƟ = -1 , so:
s = (m1·c2)2 + (m2 c2)2 + 2(E1·E2 + p1c·p2c)
Take into account that Ei >> mi·c2 and E1 ~ p1·c   and   E2 ~ p2·c , we have:
s ~ 2(E1·E2·+ E1·E2)  →  s ~ 4E1·E2
Usefull Energy: (in the Centre-of-Mass Frame = CM frame)
For the special case of identicle particles of equal momentum, colliding head-on (like the case of LHC), the C-of-M is at rest in the
lab. We have, m1 = m2 = m   and  E1 = E2 = E . Then Then equation [1] gives:
s = (m·c2)2 + (m c2)2 + 2(E·E + pc·pc)
s = 2(m·c2)2 + 2·E2 + 2(p·c)2
s = 2[(m·c2)2 +(p·c)2] + 2·E2 
but  [(m·c2)2 +(p·c)2] = E2 
so,    s = 2·E2+ 2·E2 = 4·E2

So, in the case of p-p collision at LHC, with 7 TeV per proton:
That is the energy available for new particle production in LHC collision.
For a Fixed target, considering a proton moving with energy E1 colliding with a fixed target formed for a proton at rest (m2·c2 ~ 10-
3
 TeV)  to get 14 TeV we need  E1 to be:
√s ~ √(2E1·m2·c2)
14 = √(2E1·10-3)

E1 ~ 105  TeV

It is very clear the advantage by using collider vs fixed target.


THE NUCLEAR MODELS

Any full theory of nuclear structure therefore has to contend not only with a quantum mechanical many-body problem involving
tens or, for heavier nuclei, hundreds of nucleons but also with their interaction with each other through this com­plicated potential.
An exact treatment is clearly not possible and the approach over the last half-century has been to devise increasingly sophisticated
conceptual models of the nucleus which are simple enough to enable calculations to be made, but near enough to the physical
situation to enable reasonably detailed understanding of nuclear behaviour to be achieved. Different models focus on different
aspects of nuclear behavior.

One difficulty arises from the mathematics of solving the many-body problem. If we again assume an oversimplified form for the
nuclear potential, such as a square well or an harmonic oscillator, we could in principle write down a set of coupled equations
describing the mutual interactions of the A nucleons. These equations cannot be solved analytically, but instead must be attacked
using numerical methods. A second difficulty has to do with the nature of the nuclear force itself. There is evidence to suggest that
the nucleons interact not only through mutual two-body forces, but through three-body forces as well. That is, the force on nucleon
1 not only depends on the individual positions of nucleons 2 and 3, it contains an additional contribution that arises from the
correlation of the positions of nucleons 2 and 3. Such forces have no classical analog.
NUCLEAR LIQUID DROP MODEL
In nuclear physics, liquid drop model of nucleus describes forces in atomic nuclei as if the atomic nucleus were formed by a tiny
liquid drop. But in this nuclear scale, the fluid is made of nucleons (protons and neutrons). Liquid drop model takes into account
the fact that the forces on the nucleons on the surface are different from those on nucleons on the interior where they are
completely surrounded by other attracting nucleons. This is something similar to taking account of surface tension as a
contributor to the energy of a tiny liquid drop.

Key Facts

Scattering experiments suggest that nuclei have approximately constant density.


Nuclei has its volume and surface, where forces acts differently.
In the ground state the nucleus is spherical.
If the sufficient kinetic or binding energy is added, this spherical nucleus may be distorted into a dumbbell shape and then may
be splitted into two fragments.

The Weizsaecker formula is an empirically refined form of the liquid drop model for the binding energy of nuclei. It has the
following terms:
Volume term
Surface term
Coulomb potential term
Asymmetry term
Pairing term

The liquid drop model of the nucleus, proposed by Bohr and derived by Von Weizsacker in 1935.
NUCLEAR LIQUID DROP MODEL
NUCLEAR LIQUID DROP MODEL

The Asymmetry Term.


Imagine that the neutrons and protons are in a potential
well (the nucleus).
We will assume that the well has approximately equally
spaced energy levels.
Let the spacing be ∆E.

Neutrons and protons are fermions (spin ½ particles).


So the Pauli exclusion principle allows only 2 p and 2 n in
each level.
NUCLEAR LIQUID DROP MODEL
NUCLEAR LIQUID DROP MODEL
Spin pairing in the liquid drop model
Spin pairing favours pairs of fermionic nucleons (similar to electrons in atoms) i.e. a pair with opposite spin have
lower energy than pair with same spin
Best case: even numbers of both protons and neutrons
Worst case: odd numbers of both protons and neutrons
Intermediate cases: odd number of protons, even number of neutrons or vice versa.
NUCLEAR LIQUID DROP MODEL
NUCLEAR LIQUID DROP MODEL
NUCLEAR LIQUID DROP MODEL
NUCLEAR LIQUID DROP MODEL
NUCLEAR LIQUID DROP MODEL
NUCLEAR LIQUID DROP MODEL

Binding energy per nucleon of nuclei with even mass number A.


The solid line corresponds to the Weizsacker mass formula.
Nuclei with a small number of nucleons display relatively large
deviations from the general trend, and should be considered on
an individual basis. For heavy nuclei, deviations in the form of
a somewhat stronger binding per nucleon are also observed for
certain proton and neutron numbers. This can be explained
using the so-called magic numbers.
MAGIC NUMBERS
In atomic physics, the ionization energy EI , i.e. the energy needed to
extract an electron from a neutral atom with Z electrons, displays
discontinuities around Z = 2, 10, 18, 36, 54 and 86, i.e. for noble
gases. These discontinuities are associated with closed electron shells.

An analogous phenomenon occurs in nuclear physics. There exist


many experimental indications showing that atomic nuclei possess a
shell-structure and that they can be constructed, like atoms, by filling
successive shells of an effective potential well. For example, the
nuclear analogs of atomic ionization energies are the “separation
energies” Sn and Sp which are necessary in order to extract a neutron
or a proton from a nucleus
Fig. The neutron separation energy in lead isotopes
Sn = B(Z, N ) − B(Z, N − 1) Sp = B(Z, N ) − B(Z − 1,N ) .
as a function of N . The filled dots show the
measured values and the open dots show the
These two quantities present discontinuities at special values of N or predictions of the Weizs¨acker formula.
Z, which are called magic numbers. The most commonly mentioned
are: 2 8 20 28 50 82 126.
As an example, Fig. gives the neutron separation energy of lead
isotopes (Z = 82) as a function of N . The discontinuity at the magic
number N = 126 is clearly seen.

.
MAGIC NUMBERS
It is observed experimentally that nuclei having the following values of Z and N - known as the 'magic numbers' - have various
distinctive qualities all indicating high stability:
Magic numbers
Z = 2, 8, 20, 28, 50, 82 N = 2, 8, 20, 28, 50, 82, 126
Some of the more important qualities are as follows.
1. Comparing actual nuclear masses with the smooth predictions of the semi-empirical mass formula , it is found that they are
significantly lower when Z or N is a magic number. These nuclei are therefore more tightly bound.

2. Nuclei with the above values of Z and N have significantly more stable isotopes (same Z, different N) and isotones (same N,
different Z) respectively than neighbouring nuclei. For example, for Z = 50 there are ten stable isotopes compared with an
average of around four for nearby nuclei. Similarly for N = 20 there are five stable isotones compared with an average of
around two in that region.

3. The 'doubly magic' nucleus, 4He, with Z = N = 2 is a particularly stable nucleus as are 16O with Z = N = 8 and 208
Pb (the
heaviest stable nucleus) with Z = 82 and N = 126.

4. Helium together with nuclei having N=50, 82 and 126 are particularly abundant in the universe.

5. The first excited states of nuclei with a magic number of neutrons or protons are significantly higher in energy than for
neighbouring nuclei - the nuclei are harder to disrupt. Nuclear electric quadrupole moments, are periodic as a function of Z or
N, going through zero, i.e. a compact spherical shape, at the magic numbers.
NUCLEAR SHELL MODEL
NUCLEAR SHELL MODEL

Noble gasses ionization energy


NUCLEAR SHELL MODEL
The nuclear shell model is based on the analogous model for the orbital structure of atomic electrons in atoms. Although
the liquid drop model of the nucleus has proved to be quite successful for predicting subtle variations in the mass of
nuclides with slightly different mass and atomic numbers, it avoids any mention of the internal arrangement of the
nucleons in the nucleus. We have observed that:

1. There are an abnormally high number of stable nuclides whose proton and/or neutron numbers equals the magic
numbers 2,8,20,28,50,82,126.
2. Further evidence for such magic numbers is provided by the very high binding energy of nuclei with both Z and N
being magic.
3. The abnormally high or low alpha and beta particle energies emitted by radioactive nuclei according to whether the
daughter or parent nucleus has a magic number of neutrons.
4. Nuclides with a magic number of neutrons are observed to have a relatively low probability of absorbing an extra
neutron, i.e. they have lowest of absorption cross sections for neutrons (neutron-capture cross sections).

To explain such nuclear systematics and the internal structure of the nucleus, a shell model of the nucleus has been
developed. This model uses Schrodinger‟s wave equation or quantum mechanics to describe the energetics of the
nucleons in a nucleus in a manner analogous to that used to describe the discrete energy states of electrons around the
nucleus. This model assumes:

1. Each nucleon moves independently in the nucleus uninfluenced by the motion of the other nucleons.
2. Each nucleon moves in a potential well which is constant from the center of the nucleus to its edge where it increases
rapidly by several tens of MeV.
NUCLEAR SHELL MODEL
When the model's quantum-mechanical wave equation is solved numerically, the nucleons are found to distribute themselves
into a number of energy levels.

There is a set of energy levels for protons and an independent set of levels for neutrons. Filled shells are indicated by large
gaps between adjacent energy levels and are computed to occur at the experimentally observed values of 2, 8, 20, 28, 50,
82, and 126 neutrons or protons. Such closed shells are analogous to the closed shells of orbital electrons.

However, the shell model has been useful to obtain such results that predicts the magic numbers and particularly useful in
predicting several properties of the nucleus, including:

(1) the total angular momentum of a nucleus,

(2) characteristics of isomeric transitions, which are governed by large changes in nuclear angular momentum,

(3) the characteristics of beta decay and gamma decay, and

(4) the magnetic moments of nuclei. Each nucleon moves independently in the nucleus uninfluenced by the motion of the other
nucleons.
NUCLEAR SHELL MODEL
Single-particle shell model
The basic assumption of any shell model is that despite the strong overall attraction between nucleons which provides the
binding energy, the motion of each nucleon is practically independent of that of any other nucleon. This apparent contradiction
is resolved by effects of the Pauli Exclusion Principle.
If all inter-nucleon couplings (called residual interactions) are ignored, we call the model the single-particle shell model.
In terms of Schrodinger's equation, each nucleon is then assumed to move in the same potential. The potential is spherical in
the simplest case, but there is good evidence that for nucleon numbers far from closed shells the potential should have an
ellipsoidal shape.

This model depends on two quantum numbers, the radial (total) quantum number n and the orbital quantum number l . In
nuclear physics each state is specified by n and l . Also for l = 0, 1, 2, 3, 4, 5, we use the spectroscopic letters s, p, d, f, g, h,
respectively. A state denoted by 2p therefore means that n = 2, l = 1.

The simplest useful potentials are an infinite square well potential of radius R:

or a harmonic oscillator potential:

where ω is the frequency of oscillation of the particle of mass mo. More realistic potentials are a finite square well potential
as:
NUCLEAR SHELL MODEL

Energy levels of nucleons. (a) in an infinite spherical square-well potential.


(b) in a harmonic oscillator potential.
NUCLEAR SHELL MODEL
Single-particle shell model
A simple Coulomb potential is clearly not appropriate and we need some form that describes the effective potential of all the
other nucleons. Since the strong nuclear force is short-ranged we would expect the potential to follow the form of the density
distribution of nucleons in the nucleus. For medium and heavy nuclei, the Fermi distribution fits the data and the corresponding
potential is called the Woods-Saxon form:

However, although these potentials can be shown to offer an explanation for the lowest magic numbers, they do not work for the
higher ones. This is true of all purely central potentials.
The crucial step in understanding the origin of the magic numbers was suggested that by analogy with atomic physics there
should also be a spin-orbit part, so that the total potential is:

Where L and S are the orbital and spin angular momentum operators for asingle nucleon and is an arbitrary function of the radial
coordinate. This form for the total potential is the same as that used in atomic physics except for the presence of the function .
Once we have coupling between L and S then ml and ms are no longer „good‟ quantum and we have to work with eigenstates
of the total angular momentum vector J, defined by J=L+S. Squaring this, we have:

and hence the expectation value of L·S, which is:

(We are always dealing with a single nucleon, so that s=1/2) The splitting between the two levels is thus:
NUCLEAR SHELL MODEL
Single-particle shell model
Experimentally, it is found that Vls(r) is negative, which means that
the state with j=l+1/2 has a lower energy than the state with j=l-1/2.
This is the opposite of the situation in atoms. Also, the splitting are
substantial and increase linearly with l . Hence for higher l, crossings
between levels can occur. Namely, for large l, the splitting of any two
neighboring degenerate levels can shift the j=l-1/2 state of the initial
lower level to lie above the j=l+1/2 level of the previously higher
level.

Although the spin-orbit shell model had one of the most stimulating
effects on nuclear structure physics, the simple form given
above cannot be sufficient. For example, the model cannot explain
why even-even nuclei always have a zero ground-state spin, or more
generally, why any even number of identical nucleons couples to
zero ground-state spin. Evidently there is a (residual) nucleon-
nucleon interaction which favors the pairing of nucleons with
opposing angular momenta.
Smmary on NUCLEAR SHELL MODEL
Smmary on NUCLEAR SHELL MODEL
Smmary on NUCLEAR SHELL MODEL

2 8 20 28 50 82 126.
Smmary on NUCLEAR SHELL MODEL

2 8 20 28 50 82 126.
Smmary on NUCLEAR SHELL MODEL

2 8 20 28 50 82 126.
Collective model
Three important nuclear models are the Liquid Drop Model, the Shell Model (developed by Maria Goeppert-Mayer and Hans
Jensen), which emphasizes the orbits of individual nucleons in the nucleus, and the Collective Model (developed by Aage
Bohr and Ben Mottleson), which combines both liquid drop model and shell model by including motions of the whole
nucleus such as rotations and vibrations.
For each of the liquid drop model and shell model have a specific applications, all of them succeed in the interpretation of
some phenomena and fails to explain other phenomena. So it became logical to consider each of these models is
complementary to another in a single model called the collective model as a model that combines the two models. This
model views the nucleus as having a hard core of nucleons in filled shells, as in the shell model, with outer valence nucleons
that behave like the surface molecules of a liquid drop. In addition to the successes of each of the two models, this model
has succeeded in formulating an equation to calculate the rotational energy levels to the even-even nuclei, i.e. the energy
levels of deformed nuclei are very complicated, because there is often coupling 50 between the various modes of excitation,
but nevertheless many predictions of the collective model are confirmed experimentally.

In the shell model, nuclear energy levels are calculated on the basis of a single nucleon (proton or neutron) moving in a
potential field produced by all the other nucleons. Nuclear structure and behaviour are then explained by considering single
nucleons beyond a passive nuclear core composed of paired protons and paired neutrons that fill groups of energy levels, or
shells.
In the liquid-drop model, nuclear structure and behaviour are explained on the basis of statistical contributions of all the
nucleons (much as the molecules of a spherical drop of water contribute to the overall energy and surface tension). In
the collective model, high-energy states of the nucleus and certain magnetic and electric properties are explained by the
motion of the nucleons outside the closed shells (full energy levels) combined with the motion of the paired nucleons in the
core.
Collective model
From experimental data we know that near closed shells nuclei are spherical, i.e., the equilibrium shape is a sphere. When
both the proton and neutron number differ appreciably from the magic numbers, the ground state is often found to be
axially deformed, either prolate (cigar like) or oblate (like a pancake): The nuclear Quadrupole moment Q ≠ 0
Collective model
From experimental data we know that near closed shells nuclei are spherical, i.e., the equilibrium shape is a sphere. When
both the proton and neutron number differ appreciably from the magic numbers, the ground state is often found to be
axially deformed, either prolate (cigar like) or oblate (like a pancake): The nuclear Quadrupole moment Q ≠ 0
Collective model

Bohr Aage (Ogė Bòras) 1922 06 19


- 2009 09 09, danų fizikas
teoretikas. N. Bohro sūnus. Danijos
mokslų akademijos narys (1955).

2009) 2022)
Collective model
The single-particle shell model can not properly describe the excited states of nuclei: the excitation spectra of even-
even nuclei show characteristic band structures which can be interpreted as vibrations and rotations of the nuclear
surface:

low energy excitations have a collective origin !

The liquid drop model is used for the description of collective excitations of nuclei: the interior structure, i.e., the
existence of individual nucleons, is neglected in favor of the picture of a homogeneous fluid-like nuclear matter.

The moving nuclear surface may be described quite generally by an expansion in spherical harmonics with time-
dependent shape parameters as coefficients:

where R(θ,φ,t) denotes the nuclear radius in the direction (θ,φ) at time t, and R0 is the radius of the spherical nucleus,
which is realized when all αλμ =0.

The time-dependent amplitudes αλμ(t) describe the vibrations of the nucleus with different multipolarity around the
ground state.
Collective model
The single-particle shell model can not properly describe the excited states of nuclei: the excitation spectra of even-even
nuclei show characteristic band structures which can be interpreted as vibrations and rotations of the nuclear surface:

low energy excitations have a collective origin !

VIBRATIONS ROTATIONS
Collective model

The associated excitation is the so-called breathing mode of the nucleus. Because
of the large amount of energy needed for the compression of nuclear matter, this mode is far too high in energy
to be important for the low-energy spectra discussed here. The deformation parameter  can be used to cancel
00

the overall density change present as a side effect in the other multipole deformations.

The monopole mode, is the one where the size of the nuclear fluid oscillates, i.e., where the nucleus gets compressed.
Experimentally one finds that the lowest excitation of this type, which in even-even nuclei occurs at an energy of roughly:

above the ground state. Compared to ordinary nuclear modes, which have energies of a few MeV, these are indeed high
energy modes (15 MeV for A = 216), showing the noncompressibility of the nuclear fluid. Incorporated into the average radius.
Collective model

1
R(t)  R avr  α
μ  1
1μ Y1μ (θθφ)

 R avr  α 11 Y11  α 10 Y10  α1, -1 Y1, -1


 R avr  α10 Y10 α1μ  0 for μ  0 ( α 1, -1  α11 and Y1, -1   Y11 )
3/2
1 3 
 R avr  α10   cosθ
2  2π 

The dipole mode (Figure ) by itself is not very interesting: it corresponds to an overall translation of the centre of the nuclear
fluid. One can, however, imagine a two-fluidmodel where a proton and neutron fluid oscillate against each other. This is a
collective occurs at an energy of roughly

above the ground state, close to the monopole resonance. It shows that the neutron and proton fluids stick together quite
strongly, and are hard to separate.
cannot result from action of nuclear forces (can be induced by applied EM field i.e. a photon)
Collective model

Quadrupole oscillations are the lowest order nuclear vibrational modes available for low-energy excitations in nuclei. In
almost all even-even nuclei we find a low-lying state (at excitation energy of less than 1-2 MeV ).
Collective model
Once we have created a nucleus with axial deformation, i.e., a nucleus with ellipsoidal shape, but still axial
symmetry about one axis, we can rotate the fluid around one of the non-symmetry axes to generate excitations
(Figure ). We cannot do it around a symmetry axis, since the resulting state would just be the same quantum state
as we started with, and therefore the energy cannot change. A rotated state around a non-symmetry axis is a
different quantum state, and therefore we can overlay many of these states, especially with constant rotational
velocity. This is almost like the rotation of a dumbbell, and we can predict the classical spectrum to be of the form

Where J is the classical angular momentum and is the moment of inertia for that
motion. We predict a quantum mechanical spectrum of the form

where is now the angular momentum quantum number. Naively we expect the
spectrum to be more compressed (the moment of inertial is larger) the more elongated
the nucleus becomes. It is known that certain structures in nuclei indeed describe well
deformed nuclei, up to super and hyper deformed.
In quantum mechanics, any rotation leaves the surface invariant and thus by definition
does not change the quantum-mechanical state (and energy):
•a spherical nucleus has no rotational excitations at all !
•a nucleus with axial symmetry cannot rotate around the axis of symmetry!
Nuclear Spin (I)
How do you predict the value of nuclear spin (I) based on the number of protons
and neutrons?

The shell model gives possibility to predict the spins for nucleus in ground state and even to
some exited states. The spin of completed shell is always equal to 0. Thus, for double magic
nuclei we have always I=0. The spin of nucleus is determined by outer shell (valence) nucleons.

Some examples:
Oxigen . Proton number is magic. Two lowest shells (1s and 1p) are occupied by 8 neutrons
(paired). The last neutron (9-th) is at 1d5/2 state. Thus the TOTAL angular momentum of 9-th
neutron is the spin of the nucleus I=5/2.

Nitrogen . The number of neutrons is magic. 1s shell is completed by 2 protons, 1p3/2 is comleted
by 4 protons (protons are paired and I=0). The last neutron is at 1p1/2 level. Thus, the spin of is
I=1/2.
Nuclear Spin (I)
How do you predict the value of nuclear spin (I) based on the number
of protons and neutrons?

Across the entire periodic table, nuclear spin values ranging from I = 0 to I = 8 in ½-
unit increments can be found. Protons and neutrons each have net spins of ½, but
this derives from the elementary quarks and gluons of which they are composed.
As a result of this complexity, no simple formula exists to predict I based on the
number of protons and neutrons within an atom. Nevertheless, there are some
general rules that apply to special cases:

1) Even/Even. Nuclei containing even numbers of both protons and neutrons have
I = 0 and therefore cannot undergo NMR. Examples include 4He, 12C, 16O and 32S.

2) Odd/Odd. Nuclei with odd numbers of both protons and neutrons have spin
quantumtum numbers that are positive integers. Examples include 14N (I=1), 2H
(deuterium, I=1), and 10B (I=3).

3) All others. The remaining nuclei (odd/even and even/odd) all have spins that are
half integral. Examples include 1H (I=½), 17O (I=5/2), 19F (I=½), 23Na (I=3/2), and 31P
(I=½).
Nuclear Spin (I)
In general, the most abundant isotopes of elements
with even atomic numbers all have I = 0, the only
exception being beryllium (Z=4). Its most abundant
naturally occurring isotope is 9Be, with four protons and
five neutrons, giving it a spin (I) = 3/2.

As can be seen from the periodic table above, isotopes


with integer values of spin are very uncommon. The
only nuclide with I = 2 is 204Tl, an unstable isotope with
half-life of 3.5 yrs.

Elements and isotopes with very high values of spin


(i.e., I ≥ 7) are relatively rare. Four minor isotopes of
technetium, lutetium, and rhenium all have I = 7. They
are 94Tc, 96Tc, 176Lu, and 182Re. Three isotopes have the
highest possible value for spin (i = 8): niobium (90Nb),
tantalum (181Ta), and mendelevium (258Md). The latter
(258Md) is the most abundant isotope of mendelevium,
which makes Md the "winner" for maximum spin for
common isotopes on the periodic table.
Mössbauer effect
Rudolf L. MÖSSBAUER discovers the “Recoilless Nuclear Resonance Absorption of γ-Radiation” in1958 and
receives the Nobel Prize in 1961.

Mössbauer effect: “Recoilless nuclear resonance absorption of γ-rays“


similar to Acoustic resonance between two tuning forks with same
frequency fs = fr.

Mössbauer effect: Atomic nuclei instead of tuning forks

R.L. Mössbauer made his first observation of recoilless nuclear resonant absorption in Ir!
191

R.L. Mössbauer, Z. Physik, 1958, 151, 124.


R.L. Mössbauer, Naturwissenschaften, 1958, 45, 538
Mössbauer effect
Mössbauer effect
Rudolf L. MÖSSBAUERdiscovers the “Recoilless Nuclear Resonance Absorption of γ-Radiation” in1958 and receives
the Nobel Prize in 1961.

2
E
E

R
 2
2mc
2
p
ER  2m
nuc
,
2 2
p nuc
 p, 

p 
E 
 c
2
E
E


R
2mc 2
Mössbauer effect
Mössbauer effect
An excited state (nuclear or electronic) of mean lifetime τ can never be assigned a sharp energy value, but only a value
within the energy range ∆ E, which correlates with the uncertainty in time ∆t via the Heisenberg Uncertainty Principle:
∆E∆t ≥ ħ. Weisskopf and Wigner have shown that in general:
∆E =Г = ħ/τ -> natural line width, ħ = h/2π Planck‘s constant

Transitions from an excited state (e) to the ground state (g) or vice versa involve all possible energies within the range of ∆E.
The transition probability or intensity as a function of E yields a spectral line centered around the most probable transition
energy E0.
Mössbauer effect
Mössbauer effect

Resonance absorption is observable only if the


emission and absorption lines overlap sufficiently.
Mössbauer effect
1. In optics, we have m=mass of atom, thus mc2≈109-1011eV, and Eγ =E0≈1eV  ER≈10-11-10-9eV, and typical linewidth
is Γ≈10-7eV. So, Γ>> ER and resonant absorption in optical regime is possible.

2. For gamma rays, we can analyze typical example of .


Radioactive 57Co with 270 days halflife, which may be generated in a cyclotron
and diffused into a noble metal like rhodium, serves as the gamma radiation
source for 57Fe Mössbauer spectroscopy. 57Co decays by electron capture (EC
from K-shell, thereby reducing the proton number, from 27 to 26
corresponding to 57Fe) and
initially populates the 136 keV nuclear level of 57Fe with nuclear spin quantum
number I = 5/2. This excited state decays after ~ 10 ns and populates, with 85 %
probability the 14.4 keV level by emitting 122 keV gamma quanta,
with 15 % probability the 136 keV level decays directly to the ground state of
57
Fe.
The 14.4 keV nuclear state has a halflife of ~ 100 ns. Both the halflife and the
emitted gamma quanta of 14.4 keV energy are ideally suited for 57Fe
Mössbauer spectroscopy.
Mössbauer effect

The width at half maximum of the spectral line is called the natural line width Γ
and is determined by the mean life time τ = 1.43x10-7 s for the 14.4 keV level of
57
Fe, for which Γ = h/2πτ = 4.6 x 10-9eV.

If a free atom or molecule with mass m emits or absorbs a gamma quantum of


energy Eγ, a linear momentum with recoil energy E R is imparted to the atom and
molecule, respectively, which according to the formula amounts to E R≈2x10-3 eV
and thus is six orders of magnitude larger than the transition energy E γ.

The recoil effect reduces the transition energy by ER for the emission process
and increases the transition energy by ER for the absorption process.

Thus the emission and the absorption lines are shifted away from each other by
2ER ≈ 106 Γ, and in view of the huge difference between Γ and ER an appreciable
overlap resonance is not possible. This means, that the resonant absorption of γ-
rays cannot be observed for freely moving atoms or molecules.

However in a bulk media the spectral line broadening due to Doppler effect takes
place. The line-width is: . For T=300K, (), we have . So, the resonant absorption of
γ-rays can be observed.
Mössbauer effect
At low temperatures the line-width becomes narrow, and overlapping of absorption and emission lines do not takes place. Thus one my
think that the resonant absorption is not possible. But in 1955 Mössbauer found that it takes place, because of vanished recoil.

In the solid state, crystalline or non-crystalline, recoilless emission and absorption of gamma quanta is possible, and the essentially
unshifted transition lines can (at least partially) overlap and nuclear resonance absorption can be observed. The reason is that due to the
much larger mass m<=M of a solid particle as compared to that of an atom or a molecule, the linear momentum created by emission and
absorption of a gamma quantum practically vanishes. The recoil energy caused by an emitting and absorbing atom, which is tightly
bound in the lattice, is mostly transferred to the lattice vibrational system.

In a solid (e.g. crystal) the recoil energy E R is partitioned according to ER = Etransl + Evib. Etransl refers to the linear momentum imparted to the
whole crystal of mass M. Etransl << Г, since M >>m.

Evib is converted into phonons (excitation of lattice vibrations). E vib is quantized. Within the Einstein model or Debye model: The vibrational
energy of the lattice as a whole can only change by discrete amounts 0, ħω, 2ħω, …

There is a certain probability f (known as Debye-Waller factor or Lamb-Mössbauer factor) that no lattice excitation (zero-phonon
processes) takes place during γ-emission or γ-absorption. f denotes the fraction of nuclear transitions which occur without recoil. Only for
this fraction is the Mössbauer effect observable.

Summarizing: Mössbauer effect, also called recoil-free gamma-ray resonance absorption, nuclear process permitting the 
resonance absorption of gamma rays. It is made possible by fixing atomic nuclei in the lattice of solids so that energy is not
lost in recoil during the emission and absorption of radiation. 
The Mössbauer Spectrum
Figure the schematic arrangement of a typical transmission Mössbauer experiment
(1955). The assembly consists of a source, an absorber, and a detector.

Frequency shift due to Doppler effect . Thus, for spectral line with we have resolution .
Typically, v=0.2mm/s is enough to obtain significant change of resonance conditions. Thus,
the resolution is , and for linewidths So, for , we have -> the sensitivity for spectral
measurements comparable to the natural linewidth.
Mössbauer effect: Harvard Tower Experiment
Einstein predicted a change in the energy of photons in the proximity of a gravitational field, the change being directly proportional
with the distance from the gravitational source. In the early 60’s Pound and Rebka have set to verify Einstein’s prediction. The
experiment was reprised with even higher precision by Pound and Snider
The experiment was set up in the Harvard tower. The Harvard
tower is just 22.6 meters, so the fractional gravitational red shift
between the light frequency ν at bottom of the tower and the
frequency ν0 at the top predicted by GRT given by the formula

Comparing the energy shifts on the upward and downward paths gives a predicted difference

The cleverness of the experiment lies in the fact that it sidesteps the measurement of the
frequencies at the top and of the bottom of the tower and it replaces such measurement with a very
high precision energy measurement and here is where the Mossbauer effect comes into play. The
measured difference was 5.1x10-15. The source velocity needed to compensate gravitational
frequency shift was
Nuclear reactions
Definition: A nuclear reaction is any transformation of a nucleus caused by interaction (“collision”) with an incident
particle. If a nucleus X is bombarded with a particle “a” and transforms into a nucleus Y, emitting a particle “b”, then
such a reaction is written as follows:
a X Y b or X (a, b) Y
When the nucleus collides with other nucleus or elementary particle the nuclear reaction takes place and new
particles appear. Usually the nuclear reaction are governed by nuclear force (strong interaction forces or color
forces). The exclusion is the nucleus splitting under interaction with γ-photon, dispite the fact that the of γ-quanta not
affected by nuclear interaction.
Since the nuclear forces are acting at femtometer range, the distance between interacting particles should be less
than m.
When the positively charged particles (protons, for example) are used to bombard the nuclei, the kinetic energy of
the particle should exceed the Coulomb potential (usually ~10 MeV). For negatively charged particles the barrier is
not existing and reaction takes place at the thermal energy level.
The nuclear interaction could be describe in the similar way as chemical one.
Example: . Also in truncated form .
Formula represents process where proton and neutron are knocked-out from by γ-quanta and nuleus appears.

Some examples: ,
The set of initial particles INPUT chanel, and the set of new particles is called OUPUT chanel.
Nuclear reactions
The quantative description of the nuclear reactions uses statistical methods. The reaction can be fully characterized by
effective differential cross-section per solid angle where are the azimuthual and polar angles. Actually, the dependence gives
the angular distribution of the particles. The total effective cross-section could be found by integrating on angles Usually
cross-section depends on energy of bombarding particle, target nucleus, and is in the range cm 2. (unit: 1 b (Barn) is equal to
10-24 cm²).

If de Broglie wavelength λ of matter wave of particle is λ <<R (R is nuclear radius) when the σ is defined by methods of
Explanation of cross section

Na
Nb

Nb
Na
Nuclear reactions
The quantative description of the nuclear reactions uses statistical methods. The reaction can be fully characterized by
effective differential cross-section per solid angle where are azimuthual and polar angles. Actually, the dependence gives the
angular distribution of the particles. The total effective cross-section could be found by integrating on angles Usually cross-
section depends on energy of bombarding particle, target nucleus, and is in the range cm 2. (unit: 1 b (Barn) is equal to 10-
24
 cm²).

If de Broglie wavelength λ of matter wave of particle is λ <<R (R is nuclear radius) when the σ is defined by methods of
Nuclear reactions
If de Broglie wavelength λ of matter wave of particle is λ <<R (R is nuclear radius) when the σ is defined by methods of
geometrical optics. In such case reactions are non-resonant and nuclei are not transparent, thus
For slow particles λ ≥ R and the methods of geometrical optics are not valid. The methods wave optics should be used. Thus the is
defined by λ and not by R.
Several conservation of such quantities as energy, momentum, angular momentum, electric charge, lepton number, baryon number
end etc. should be valid. This makes some restrictions telling us which reactions are possible or not.
Let’s take a look on energy conservation. For relativistic particle (reaction ) we can write:

We also can write:

If Q > 0 the rest mass is decreased, and Q is the energy (as kinetic energy of particles) released during the interaction, and could
be included in equation:

Just means that if particles (a, A) partcipate in reaction being at rest, the particles are moving with the total kinetic
energy Q.
If Q > 0, the reaction called exothermic. Example: +17MeV.
If Q < 0, the reaction called endothermic. Example: +17MeV (alfa particles needed with 17 MeV kinetic energy for
creation Li atom and proton at rest)
Nuclear reactions:compound-nucleus model
Compound-nucleus model (tarpinio branduolio modelis). The compound nucleus is the intermediate state formed in a
nucleus reaction. It is normally one of the excited states of the nucleus formed by the combination of the incident particle and
target nucleus. The compound nucleus is excited by both the kinetic energy of the projectile and by the binding nuclear energy.
To understand the nature of nuclear reactions, the classification according to the time scale of these reactions has to be
introduced. Interaction time is critical for defining the reaction mechanism.
There are two extreme scenarios for nuclear reactions:
• A projectile and a target nucleus are within the range of nuclear forces for a very short time allowing for an interaction of a
single nucleon only. These types of reactions are called direct nuclear reactions.
• A projectile and a target nucleus are within the range of nuclear forces, allowing for a large number of interactions
between nucleons. These types of reactions are called the compound nucleus reactions.
Basic characteristics of direct reactions:
• The direct reactions are fast and involve a single-nucleon interaction.
• The interaction time must be very short (~10-22 s).
• The direct reactions require incident particle energy larger than ∼ 5 MeV/Ap. (Ap is the atomic mass number of a
projectile)
• Incident particles interact on the surface of a target nucleus rather than in the volume of a target nucleus.
• Products of the direct reactions are not distributed isotropically in angle, but they are forward-focused.
• Direct reactions are of importance in measurements of nuclear structure.
Nuclear reactions:compound-nucleus model
The compound nucleus is the intermediate state formed in a compound nucleus reaction. It is normally one of the excited states of the
nucleus formed by the combination of the incident particle and target nucleus. Suppose a target nucleus X is bombarded with particles a.
In that case, it is sometimes observed that the ensuing nuclear reaction occurs with appreciable probability only if the particle’s energy is
in the neighborhood of certain definite energy values. These energy values are referred to as resonance energies. The compound nuclei of
these certain energies are referred to as nuclear resonances. Resonances are usually found only at relatively low energies of the
projectile. The widths of the resonances increase in general with increasing energies. At higher energies, the widths may reach the order of
the distances between resonances, and then no resonances can be observed. The narrowest resonances are usually the compound states of
heavy nuclei (such as fissionable nuclei) and thermal neutrons (usually in (n,γ) capture reactions). The observation of resonances is by no
means restricted to neutron nuclear reactions
Basic characteristics of compound nucleus reactions:
• The compound nucleus is a relatively long-lived intermediate state of the particle-target composite system.
• The compound nucleus reactions involve many nucleon-nucleon interactions.
• A large number of collisions between the nucleons leads to a thermal equilibrium inside the compound nucleus.
• The time scale of compound nucleus reactions is of the order of 10-18 s – 10-16 s.
• The compound nucleus reactions are usually created if the projectile has low energy.
• Incident particles interact in the volume of a target nucleus.
• Products of the compound nucleus reactions are distributed near isotropically in angle (the nucleus loses memory of how it was
created – Bohr’s hypothesis of independence).
• The decay mode of the compound nucleus does not depend on how the compound nucleus is formed.
• Resonances in the cross-section are typical for the compound nucleus reaction.
Nuclear reactions:compound-nucleus model
• The compound nucleus model (the idea of compound nucleus formation) was introduced by Danish physicist Niels Bohr in
1936. This model assumes that the incident particle and the target nucleus become indistinguishable after the collision and
constitute the nucleus’s particular excited state – the compound nucleus. The projectile has to suffer collisions with
constituent nucleons of the target nucleus until it has lost its incident energy to become indistinguishable. Many so these
collisions lead to a complete thermal equilibrium inside the compound nucleus. The compound nucleus is excited by both
the kinetic energy of the projectile and by the binding nuclear energy.
• This compound system is a relatively long-lived intermediate state of the particle-target composite system. From the
definition, the compound nucleus must live for at least several times longer than is the time of transit of an incident particle
across the nucleus (~10-22 s). The time scale of compound nucleus reactions is of the order of 10 -18 s – 10-16 s, but lifetimes as
long as 10-14 s have also been observed.
• A very important feature and a direct consequence of the thermal equilibrium inside a compound nucleus is that the mode of
decay of the compound nucleus does not depend on how the compound nucleus is formed. Many collisions between
nucleons lead to the loss of information on the entrance channel from the system. The decay mechanism (channel) that
dominates the decay of C* is determined by the excitation energy in C*.
Nuclear reactions:compound-nucleus model
These reactions can be considered as two-stage processes.
• The first stage is the formation of a compound nucleus expressed by σa+X➝C*
• The second stage is the decay of a compound nucleus expressed by PC*➝b+Y
• The result cross-section of certain reaction a+X➝[C*]➝b+Y is given by σ(a,b)= σa+X➝C* . PC*➝b+Y

For the compound nucleus, peaks in the cross-section are typical. Each peak manifests a particular compound state of the nucleus. These
peaks and the associated compound nuclei are usually called resonances. The behavior of the cross-section between two resonances is
usually strongly affected by the effect of nearby resonances.
Resonances (particular compound states) are mostly created in 
neutron nuclear reactions, but it is by no means restricted to neutron nuclear reactions.
The quantum nature of nuclear forces causes the formation of resonances. Each nuclear
reaction is a transition between different quantum discrete states or energy levels. The
discrete nature of energy transitions plays a key role. Suppose the energy of the
projectile (the sum of the Q value and the kinetic energy of the projectile) and the
energy of the target nucleus are equal to a compound nucleus at one of the excitation
states. In that case, a resonance can be created, and a peak occurs in the cross-section.
The allowable state density in this energy region is much lower for the light nucleus,
and the “distance” between states is higher. For heavy nuclei, such as 238U, we can
observe a large resonance region in the neutron absorption cross-section.
Direct Nuclear Reactions
Nuclear reactions that occur in a time comparable to the time of transit of an incident particle across the nucleus (~10-22 s) are
called direct reactions. Interaction time is critical for defining the reaction mechanism. The very short interaction time allows
for an interaction of a single nucleon only (in extreme cases). There is always some non-direct (multiple internuclear
interactions) component in all reactions, but the direct reactions have this component limited. The reaction has to occur at high
energy to limit the time available for multiple internuclear interactions.
Direct reactions have another very important property. Products of a direct reaction are not distributed isotropically in angle,
but they are forward-focused. This reflects that the projectiles make only one, or very few, collisions with nucleons in the target
nucleus, and its forward momentum is not transferred to an entire compound state.
The cross-sections for direct reactions vary smoothly and slowly with energy in contrast to the compound nucleus reactions.
These cross-sections are comparable to the geometrical cross-sections of target nuclei.
Types of direct reactions:
• Elastic scattering in which a passing particle and targes stay in their ground states.  
• Inelastic scattering in which a passing particle changes its energy state.  For example, the (p, p’) reaction.
• Transfer reactions in which one or more nucleons are transferred to the other nucleus. These reactions are further classified
as:
• Stripping reaction in which one or more nucleons are transferred to a target nucleus from passing particles. For example,
the neutron stripping in the (d, p) reaction.
• Pick-up reaction in which one or more nucleons are transferred from a target nucleus to a passing particle. For example,
the neutron pick-up in the (p, d) reaction.
• Break-up reaction in which a breakup of a projectile into two or more fragments occurs.
• Knock-out reaction in which a single nucleon or a light cluster is removed from the projectile by a collision with the target.
Direct Nuclear Reactions

Example: This threshold reaction of a fast neutron with an


isotope 10B is one of the ways how radioactive is generated.
Direct Nuclear Reactions
Nuclear reactions that occur in a time comparable to the time of transit of an incident particle across the nucleus (~10-22 s) are
called direct reactions. Interaction time is critical for defining the reaction mechanism. The very short interaction time allows
for an interaction of a single nucleon only (in extreme cases). There is always some non-direct (multiple internuclear
interactions) component in all reactions, but the direct reactions have this component limited. The reaction has to occur at high
energy to limit the time available for multiple internuclear interactions.
Direct reactions have another very important property. Products of a direct reaction are not distributed isotropically in angle,
but they are forward-focused. This reflects that the projectiles make only one, or very few, collisions with nucleons in the target
nucleus, and its forward momentum is not transferred to an entire compound state.
The cross-sections for direct reactions vary smoothly and slowly with energy in contrast to the compound nucleus reactions.
These cross-sections are comparable to the geometrical cross-sections of target nuclei.
Types of direct reactions:
• Elastic scattering in which a passing particle and targes stay in their ground states.  
• Inelastic scattering in which a passing particle changes its energy state.  For example, the (p, p’) reaction.
• Transfer reactions in which one or more nucleons are transferred to the other nucleus. These reactions are further classified
as:
• Stripping reaction in which one or more nucleons are transferred to a target nucleus from passing particles. For example,
the neutron stripping in the (d, p) reaction.
• Pick-up reaction in which one or more nucleons are transferred from a target nucleus to a passing particle. For example,
the neutron pick-up in the (p, d) reaction.
• Break-up reaction in which a breakup of a projectile into two or more fragments occurs.
• Knock-out reaction in which a single nucleon or a light cluster is removed from the projectile by a collision with the target.
Copyright © 2010 Pearson Education,
Chapter Eleven 196
Inc.
Nuclear reactors
• www.youtube.com/watch?v=uYrhWO_ZLYw

• www.youtube.com/watch?v=ekub_xEiUww

You might also like