(De Gruyter Textbook) Frank Rösch - Nuclear - and Radiochemistry, Volume 1 - Introduction-De Gruyter (2022)
(De Gruyter Textbook) Frank Rösch - Nuclear - and Radiochemistry, Volume 1 - Introduction-De Gruyter (2022)
(De Gruyter Textbook) Frank Rösch - Nuclear - and Radiochemistry, Volume 1 - Introduction-De Gruyter (2022)
Macromolecular Chemistry.
Natural and Synthetic Polymers
Mohamed Elzagheid,
ISBN ----, e-ISBN ----,
e-ISBN (EPUB) ----
Bioanalytical Chemistry.
From Biomolecular Recognition to Nanobiosensing
Paolo Ugo, Pietro Marafini, Marta Meneghello,
ISBN ----, e-ISBN ----,
e-ISBN (EPUB) ----
Quantum Chemistry.
An Introduction
Michael Springborg, Meijuan Zhou,
ISBN ----, e-ISBN ----,
e-ISBN (EPUB) ----
Green Chemistry.
Principles and Designing of Green Synthesis
Syed Kazim Moosvi, Waseem Gulzar Naqash, Mohd. Hanief Najar,
ISBN ----, e-ISBN ----,
e-ISBN (EPUB) ----
Frank Rösch
Nuclear- and
Radiochemistry
Volume 1: Introduction
ISBN 978-3-11-074271-8
e-ISBN (PDF) 978-3-11-074272-5
e-ISBN (EPUB) 978-3-11-074280-0
www.degruyter.com
Preface
Except for the element hydrogen with its most common stable isotope 1H, the nuclei
of all atoms consist of mixtures of protons and neutrons.1 For certain well-defined
nucleon compositions (i.e. protons + neutrons), nuclei are stable – they exist forever.
This holds true for one or more nuclei of the isotopes of almost all chemical elements
(except technetium, Z = 43) ranging from hydrogen (Z = 1) to bismuth (Z = 83). Yet,
there are less than 300 stable nuclei altogether.
In contrast, all the known elements beyond bismuth (i.e. from Z = 84 to Z = 119)
and many of the lighter elements (i.e. from Z = 1 to Z = 83) have isotopes comprising
nuclei of nucleon compositions not suitable to stability. This means that there are
more than 3000 unstable nuclei that must transform into new nuclei of more stable
nucleon compositions. This textbook is about the many atomic nuclei and their
characteristic states that form the atoms composing our physical world: stable nu-
clei (K), nuclei with excited nuclear levels (⊙K), unstable nuclei (✶K), nuclei at
ground state (gK), or metastable states (mK).2
Radiochemical transformations proceed in an exothermic way. The velocity of a
transformation is expressed by a transformation constant and/or half-life.
The transformation itself is accompanied by the release of various kinds of
emissions. It is the velocities and emissions of these transformations that stand for
the phenomenon of “radioactivity”.
The understanding of basic nuclear and radiochemical processes is a prerequi-
site to exploring the potential of radionuclides and radioactivity as a source of en-
ergy, as tools in fundamental research and analytics, for environmental purposes,
for industrial applications, in medicine, etc. For example, nuclear fission is one of
the most important sources of electricity, while the beta- and alpha-decay processes
are the essence of several molecular imaging processes adopted in diagnostic nu-
clear medicine and for patient treatments, respectively. These and several other
topics are addressed in the second volume of this textbook.
The first volume of this textbook introduces the basic aspects of these pro-
cesses. It focuses on explaining the fundamentals, rather than on specific details
and mathematical treatments. Mathematics, and in particular (quantum) physical
models, are referred to only when conventional physics cannot explain an impor-
tant experimental observation.
Thus, this textbook serves as a guide to qualitatively understand the essence of
radioactivity, considering questions such as:
https://doi.org/10.1515/9783110742725-202
VI Preface
https://doi.org/10.1515/9783110742725-203
Contents
Preface V
9 α-Emission 239
9.1 Introduction 239
9.2 Mass balances in α-transformations 241
9.3 Pathways of α-emission 243
9.4 Energetics 246
9.5 Velocities of α-transformation 252
9.6 Quantum mechanics of α-transformation phenomena 256
9.7 Excited states 260
9.8 Example applications 262
9.9 Outlook 269
Contents XI
14 Appendix 433
Index 451
1 The atom’s structure I: Electrons and shells
Aim: The concept of atoms as the smallest, indivisible constituents of matter is in-
troduced. It arises from more than two millennia old Greek philosophy, reaches the
atom’s renaissance in the nineteenth century, and reflects the dramatic improve-
ments achieved at the beginning of the twentieth century.
Atoms are made of electrons, which exist in the shell of an atom, and a set of
different particles located in the atom’s nucleus. To understand the chemical prop-
erties of an atom, the electron shell structure – i.e. the number, characteristics, and
transitions of electrons in various shells – is essential. To a large extent, this under-
standing needs a significant reflection on quantum physics.
Developments of concepts of ancient and modern atom theory are introduced,
turning from the atomic philosophy of being the ultimate particle to concepts of
the atom being a substance of various, subatomic particles. The latter ideas are il-
lustrated as a scientific need to understand pioneering experiments in chemistry
and physics, i.e. to correlate experimental evidence and subsequent theoretical
explanation.
The latter is meant to serve as an introduction to (quite similar) considerations
relevant to the structure of nucleons in the atomic nucleus.
The concept of an atom was approached about 24 centuries ago when ARISTOTLE
(384–322 BC) suggested that all existing matter is built of four components: air,
water, fire, and earth. PLATO (427–347 BC) added a fifth element, the ether, which
allows for interaction and transformation between the others. The five Platonic sol-
ids are illustrated in Fig. 1.1. Interestingly, they are each composed of just one of
only three single abstract geometric figures (equilateral triangle, square, and a reg-
ular pentagon), which when assembled form the tetrahedron, hexahedron, octahe-
dron, icosahedron, and dodecahedron.
At almost the same time, the Greek philosopher LEUCIPPUS (ca. 450–370 BC) and
his scholar DEMOCRITUS (460–371 BC) proposed a material world made of indivisible
components. Assuming that every material could be divided into smaller parts, and
these again into even smaller fragments; finally, a level of indivisible things is
reached. They called them atomos, which is the Greek name of atoms, the indivisi-
ble. The philosophers ascribed them with defined properties, namely to be:
– very small,
– and thus invisible,
https://doi.org/10.1515/9783110742725-001
2 1 The atom’s structure I: Electrons and shells
– indivisible,
– hard,
– of different forms (however without color, taste, or smell),
– moving spontaneously and continuously (in an empty space, i.e. in vacuo or in
“ether”).
Fig. 1.1: The five Platonic solids. Each is composed of a number of identical isosceles surfaces:
tetrahedron (4 triangles), hexahedron (6 squares), octahedron (8 triangles), icosahedron (20
triangles), dodecahedron (12 pentagons).
To conceive of the ultimate building blocks1 as having “different forms” included the
elegant idea that the matter built of them finally reveals properties of the “atoms”,
which themselves remain invisible. Ultimately, the building blocks are not only the
composites of our material world, they are responsible for its properties.
It took more than 22 centuries until this conceptual idea was proven experimentally.
Chemists like LAVOISIER (1743–1794), PROUST (1755–1826), and DALTON (1766–1844)
realized that individual chemical elements combine mass-wise to form compounds
according to a set of rules.
It would become obvious only more than 2000 years later that this is not really true; it is more
the number and the internal infrastructure of a well-defined, but small group of subatomic building
blocks called “elementary particles” which define an “atom”. These constructs indeed define the
physical and chemical properties of atoms and chemical elements.
1.1 The philosophy of atoms 3
Tab. 1.1: The laws of conservation of mass and of definite proportions exemplified for the formation
of water out of the two gases hydrogen and oxygen. The ratio between the two elements is 1:8 if
masses are counted, and 2:1 if moles are considered. Using the AVOGADRO number to convert moles
into numbers of atoms, it is obvious that the overall number of atoms of the reactants (H2 and O2) is
the same for the product (H2O), and that the number of both hydrogen and oxygen atoms starting the
reaction did not change, even if the reaction product is chemically a completely different species.
H + O ! HO
g = g
H × .⋅+ = × .⋅+
O × .⋅+ = × .⋅+
Within these reactions, atoms do not disappear, i.e. are neither destroyed nor
created.
DALTON’S key point was the renaissance of the assumptions made by the Greek
philosophers more than two millennia ago. Chemical elements are made of atoms,
representing the smallest indivisible constituents of matter (Fig. 1.2).
The identification of electricity and electrolysis were key factors in the development
to make the “classical” atom “transparent”. These two effects appeared to be
completely new to science and stimulated new theories on the atom. FARADAY
(1791–1867) experimentally observed that electric current may disrupt chemical
compounds. STONEY (1826–1911) suggested the existence of “carriers of electric
charges”, and that those carriers were associated with the atom. In 1891 he called
them “electrons”. PLÜCKER (1801–1868) tried to conduct electric current through
vacuum and in 1859 observed cathode rays. THOMSON (1856–1940) systematically
analyzed these rays and realized experimentally that small, electrically-charged
particles – electrons – were emitted from the atoms of the cathode (see Fig. 1.3).
Independent of the type of hot cathode, the electron properties were the same.
1.2 The inner structure of the atom 5
ELEMENTS
Simple
1 2 3 4 5 6 7 8
9 10 11 12 13 14 15 16
I Z C L
17 18 19 20
S P G
Binary
21 22 23 24 25
Ternary
26 27 28 29
Quaternary
30 31 32 33
Septenary
36 37
Fig. 1.2: DALTON’S “New System of Chemical Philosophy” describing the elements with individual
symbols and chemical compounds as individual elements interconnected in specific relationships
(stoichiometry), without losing their identity.
This was the experimental evidence of the existence of the electron. THOMSON2
even managed to measure the intensity of the electrons’ deviation in external electric
and magnetic fields. From these observations, he derived the first experimental value
for the charge-to-mass quotient of the electron, which today is q/m = − 1.759⋅108 C/g.
JJ THOMSON received the Nobel Prize in Physics for this fundamental invention in 1906 “in recog-
nition of the great merits of his theoretical and experimental investigations on the conduction of
electricity by gases”.
6 1 The atom’s structure I: Electrons and shells
partially evacuated
glass tube
– +
HV
(a)
+ –
HV
Fig. 1.3: Gas discharge tube experiments. Electric current (HV = high voltage) conducted through
vacuum causes emissions – the electrons (a). Gas discharge of hydrogen gas causes a flux of
“naked” protons towards a fluorescent screen (b).
The most striking point to realize is that the electron itself is a subatomic parti-
cle, thus challenging DEMOCRITUS’S and DALTON’S concept of the atom as the ulti-
mate indivisible species.
Further experiments with similar devices, namely the gas-filled discharge tube,
soon revealed another form of particle emission. In 1886 GOLDSTEIN induced gas dis-
charge and detected positively charged components, which were emitted through a
hole inside the cathode (see Fig. 1.3). Similar to JJ THOMSON’S approach, the degree
of deviation of these positively charged ions of the gas (cations) by external electric
and magnetic fields could be analyzed. On the basis of hydrogen emissions, W
WIEN derived in 1897 a q/m value of + 9.58⋅104 C/g for this cation. He named it “pro-
ton”3 as a tribute to the ancient Greeks; proton means the first, kind of primordial
particle: another subatomic particle.
Nobel Prize in Physics, 1911, “for his discoveries regarding the laws governing the radiation of
heat”.
1.3 The shell and the nucleus 7
With the discussion of two types of subatomic particles, a new conceptual level was
reached. However, further insights into the atom’s structure were not possible with
the classic infrastructure of chemical and physical experimental and analytical
methodologies available in the nineteenth century.4 Similar to the impact of the ap-
pearance of electricity on the progress of atomic theory, it was the phenomenon of
radioactivity that opened the door to the experimental studies ahead. (Interestingly,
while radioactivity required significant progress in the understanding of the inner
structure of the atom, radioactivity itself was at the same time a tool to do so.)
In 1896 WC ROENTGEN first measured a new type of radiation originating from
uranium ores, which he was unable to identify and therefore named X-rays.5 Soon
after, M CURIE chemically separated the main elements from uranium ore responsi-
ble for X-ray emission: radium and (later) polonium.6 Samples of radioactive ra-
dium became a popular tool for a variety of seminal experiments – not only at M
and P CURIE’S laboratory in Paris, but also elsewhere. By performing experiments
similar to those conducted for the electron and the proton, it was discovered that
the principle emissions were small, positively charged species.7 The charge was
twice that of the proton and the mass fourfold that of the proton. Since they were
the first radioactive particle emission investigated, they were named “α-particles”
after the first letter of the Greek alphabet – in particular by E RUTHERFORD.
As we know today, excursions into the atom required subatomic tools and energies much higher
than those provided by classic chemistry and physics.
(First) Nobel Prize in Physics, 1901, “in recognition of the extraordinary services he has rendered
by the discovery of the remarkable rays subsequently named after him”.
Nobel Prize in Physics, 1903, “in recognition of the extraordinary services they (i.e. with Pierre
CURIE) have rendered by their joint researches on the radiation phenomena discovered by Professor
Henri BECQUEREL; and Nobel Prize in Chemistry, 1911”, “in recognition of her services to the ad-
vancement of chemistry by the discovery of the elements radium and polonium, by the isolation of
radium and the study of the nature and compounds of this remarkable element”.
See Chapter 5 for the radioactive transformations originating from uranium, radium etc. and
Chapter 9 for details on α-transformations.
8 1 The atom’s structure I: Electrons and shells
scattered
α-particles non-scattered
α-particles
gold atom
beam of (representing
α-particles thin Au foil)
α-source
Fig. 1.4: Sketch of the RUTHERFORD experiment, analyzing α-particle8 scattering9 on atoms
of a thin gold foil.
This observation was not in agreement with the THOMPSON model of an atom, which
considered electrons and protons to be distributed homogeneously. In this case, the
α-particles should have been unable to penetrate the gold atom for one of two rea-
sons: the density of the 76 protons and the 76 electrons inside the gold atom, or
because of coulombic repulsion between the 76 protons and the incoming +2 charged
α-particles. Alternatively, if the α-particles could penetrate this substance, it should
have occurred along their line of origin. RUTHERFORD explained the experimental re-
sult by suggesting a new “RUTHERFORD atom model”. It divided the atom into two
The α-particles emitted from the source originate from the 226Ra itself, but also from other radio-
nuclides involved in 226Ra transformation processes such as 218Po and 214Po.
Despite penetrating the atom and scattering (elastically or inelastically), another option would
be a nuclear reaction (see Chapter 13). This, however, could not happen in RUTHERFORD’S experi-
ment because of the kinetic energies of the α-particles relative to the high proton number of the
gold nucleus.
1.3 The shell and the nucleus 9
principal components – the protons, concentrated inside a nucleus, and the elec-
trons, located within a shell.
RUTHERFORD’S model created a sufficiently empty space between the two sub-
atomic particles known at that time. This space was large enough for two processes:
to permit α-particles entering the gold atom just between the nucleus and the outer
shell to pass through (and reach the detectors behind the gold foil), and at the
same time to reflect and/or scatter incoming α-particles which hit the nucleus (thus
reaching detectors in a 4π geometry positioned around the gold foil).
According to the RUTHERFORD atom model,10 each chemical element is character-
ized by its proton number (Z). For a neutral atom, the number of protons within the
nucleus equals the number of electrons in its shells. Electrons positioned in a shell
were considered to “circulate” around the nucleus. Significant energy was antici-
pated for the speed of this movement in order to withstand the coulomb attraction of
the protons and prevent the electrons from being sucked into the oppositely-charged
nucleus (Fig. 1.5).
Fig. 1.5: Schematic representation of the successive levels of atomic “philosophy”: (a) The atom as
an indivisible component of matter (Democritus, Dalton). (b) Subatomic structures of electrons and
protons distributed homogeneously inside the atom (Thomson). (c) Electrons and protons
positioned in the shell and nucleus of the atom, respectively (Rutherford). Sizes are not to scale:
the proton mass is about 2000× that of the electron mass, with which size scales accordingly. The
diameter of a spherical nucleus (<10 fm) is about 1000× smaller than the diameter of the whole
atom (ca. 100 pm or 1 Å). Consequently, the “empty” space within the electron shell and the
nucleus is much larger than illustrated. The area of the gold atom is on the order of 10−20m2, that
of the gold nucleus about 10−28m2, making an area of empty space of about 99.999999% of the
atom’s cross-section (see Chapter 2). About 99.99% of the atom’s mass is concentrated within the
little nucleus!
Nobel Prize in Chemistry, 1908, “for his investigations into the disintegration of the elements,
and the chemistry of radioactive substances”.
10 1 The atom’s structure I: Electrons and shells
While the new atom model could explain the experimentally observed α-scattering,
it provoked fundamental questions. Three main concerns were:
1. Why should all the protons be “willing” to concentrate tightly together and
stick in the little nucleus – despite the coulomb repulsion of the positively
charged particles?
2. When a chemical element is identified by its number of protons (and the same
number of electrons in the case of a neutral atom) – why may the chemical
properties of one and the same element differ for different chemical compounds
containing the same atom?
3. Why might electrons circulate in the shell? Electrically-charged moving par-
ticles are known through classic electrodynamics to lose energy. Electrons cir-
culating in a shell around the nucleus and within the field of the nucleus must
emit energy. Consequently, they are expected to approach closer and closer to
the nucleus – until they are finally trapped there. The RUTHERFORD atom thus
would be not stable, i.e. could not exist!
Although this idea was deduced somehow from the PTOLEMAIC conception of our
planetary system – with electrons circulating around the atomic nucleus like plan-
ets around our sun – this new “BOHR atom model”11 was not in accordance with
classical physics. However, it was able to perfectly explain new experimental obser-
vations obtained for the line spectrum of hydrogen (see Fig. 1.6).
Once the hydrogen atom received energy, e.g. by elevated temperature, a set of
individual photon emissions was observed, lying within the wavelength spectrum
Nobel Prize in Physics, 1922, “for his services in the investigation of the structure of atoms and
of the radiation emanating from them”. Remarkably, AAGE NIELS BOHR, son of NIELS BOHR, continued
research on atomic nuclei and in particular on the interaction among nucleons causing nuclear dis-
tortion. He shared the Nobel Prize in Physics, 1975, “for the discovery of the connection between
collective motion and particle motion in atomic nuclei and the development of the theory of the
structure of the atomic nucleus based on this connection”.
1.3 The shell and the nucleus 11
∞
7
6
(H ) (H ) (H ) (H )
=410.1 nm =434.0 nm =486.1 nm =656.3 nm
Fig. 1.6: Electron transitions of the hydrogen emission spectrum (schematic). Each emission series
(Lyman, Balmer, etc.) corresponds to a different final energy level (n). Notably, emissions in the
BALMER series occur at visible wavelengths, shown below in nm.
accessible to the human eye. However, there was just one electron within the neu-
tral hydrogen atom – but sets of characteristic emissions.
This observation is explained by allowing the electron to occupy several shells (n)
of distinct energy. The hydrogen electron is assumed to normally exist in the ener-
getically lowest shell (n = 1). When energetically stimulated, it may reach an orbital
shells of higher energy (n = 2, 3, . . . ), i.e. further from the hydrogen nucleus. Imme-
diately, it would fall back to a lower-energy shell.
Once a certain shell is taken as the baseline, the electron is energetically excited
to reach a higher-energy shell, from where it immediately returns. The difference in
12 1 The atom’s structure I: Electrons and shells
1 Z2
EBðeÞ ðnÞ = − constant · 2
= RH · 2 (1:1)
n n
n= 1 2 3 4
Fig. 1.7: Schematic energy scale of electron binding energy in proportions according to eq. (1.1).
Starting with the energy level for n = 1, any higher energy level is fractionated due to the EB(e)(n = 1)
value as 1/4, 1/9, 1/16, 1/25 etc. for EB(e)(n = 2), EB(e)(n = 3), EB(e)(n = 4), EB(e)(n = 5), respectively.
For different atoms, electron binding energies in individual shells are very much
dependent on the number of protons in the nucleus: the higher Z is, the higher EB(e)
(n) is by ≈(1/n)2. With eq. (1.1), it is straightforward to calculate the electron binding
energy for any electron shell of any atom. For gold (Z = 76), the n = 1 electron bind-
ing energy is − 78 553.6 eV; for shells 2, 3, 4 and 5 it is − 19 638.4 eV, − 8 728.2 eV,
− 4 909.6 eV and − 3 142.1 eV, respectively. The higher Z is and the closer the elec-
tron is to the nucleus, the higher the energy needed to remove it from its shell. For
example, the K-shell electron of gold is bound almost six thousand times stronger
than that of hydrogen.13
In atomic physics, the energy unit electron volt (eV) is preferred to joule (J): 1 eV = 1.602177⋅10−19
J; 1 J = 1 kg⋅m2 / s2. Alternatively, energy might be translated into mass via E = mc2, and the RYDBERG
constant may be expressed in terms of mass, typically related to the mass of an electron or to the
atomic unit.
For the following chapters on binding energies of protons and neutrons inside the atom nu-
cleus, it is interesting to note this typical range of electron binding energy, namely less than
100 keV. Typically, chemical reactions proceed at eV scales: 1–7 eV between identical atoms (atom
1.3 The shell and the nucleus 13
The 1/22 term corresponds to the n = 2 baseline (BALMER); for the other series, this
would become 1/12 (LYMAN), 1/32 (PASCHEN), 1/42 (BRACKETT), etc., with n always corre-
sponding to the higher shell level. The lowest energy emission of the BALMER series
(α-line) thus corresponds to the EB(e)(n = 3) ! EB(e)(n = 2) transition, and according to
eq. (1.2) the wavelength λ of this quantum emission is 656 nm. It represents a photon
lying in the red area of the optical spectrum. The whole BALMER series is within the
visible part of the electromagnetic spectrum.14 Figure 1.8 illustrates the transitions
within the different main electron shells of hydrogen, indicating the individual ΔEB(e)(n)
values for the BALMER series (see also Fig. 1.6).
The energy of the electromagnetic radiation is found using E = m⋅c2 and E = h⋅υ
(frequency υ, wavelength λ, and wavenumber ν̅). The PLANCK constant h was intro-
duced in quantum mechanics to identify the sizes of energy quanta.15 It is basically the
proportionality constant between the energy E of a photon and its frequency υ (E = hυ).
Because frequency is proportional to wavelength by a proportionality factor c (the
speed of light), the PLANCK constant also reflects the correlation E = h⋅c/υ.
ΔE=h·ν (1:3)
The BOHR atom model not only attributed electrons to specific “allowed” orbital
shells, it also arranged the number of electrons to fit into individual shells. A crude
binding), 3–15 eV for ion binding, 1–8 eV for metal binding and 0.01–0.2 eV for VAN DER WAALS
interactions. The binding energies of nucleons will be within an MeV scale.
The line spectrum of hydrogen is a “brilliant” tool in astrophysics, for example. Detecting the α-
line and the deviation caused by the DOPPLER effect derives the speed of stars moving in the
universe.
Units of h are J⋅s or eV⋅s, i.e. 6.626 ⋅ 10−34 J⋅s or 4.135⋅10−15 eV⋅s.
14 1 The atom’s structure I: Electrons and shells
0 eV
n=7 (Q)
n=6 (P)
n=5 (O)
n=4 (N)
n=3 (M)
430.0
656.3
486.1
410.1
n= 1 2 3 4 ...
proton
n=2 (L) (= nucleus of
hydrogen atom)
Fig. 1.8: Electronic transitions of the BALMER series. Four emission wavelengths are shown,
including the α-line at 656.3 nm in red. The binding energy of the “normal” n = 1 electron is
13.6 eV.
viewpoint is to believe that shells close to the nucleus are “smaller” than those more
distant from the nucleus. “Larger” shells can accept more electrons. Figure 1.9 illus-
trates the filling of electrons into the first three shells of an atom. Shell n = 1 is consid-
ered “closed” when it contains 2 electrons, shell n = 2 accepts a maximum of 8
electrons, shell n = 3 accepts 18, etc.
Over the centuries, a variety of individual chemical elements have been identified.
Their physical and chemical properties were measured according to the analytical
technologies available. This created a huge dataset of element-specific chemical
and physical parameters. It appeared that some elements behaved chemically quite
similarly and could thus be considered group-wise. Chemists like MENDELEEV (in
1896) realized that there should be a system to classify and group the elements ac-
cordingly. An initial criterion was to line up the elements following their increasing
mass, starting with the lightest one, hydrogen, with number Z = 1. The other crite-
rion was to consider the chemical similarities of some elements and to put them
into different groups.
1.3 The shell and the nucleus 15
ELECTRON
MAIN
SHELL
NUMBER
n=4
n=3
n=2
n=1
ELECTRON
ORBITAL
SHELL
NUMBER
l=1
l=0
NUCLEUS
Fig. 1.9: An atom of the element calcium in its neutral atomic state containing 20 protons and 20
electrons. All electrons are distributed within the various main shells n according to the maximum
shell occupancy numbers and electron binding energies (qualitatively).
Fig. 1.10: Periodic Table of Elements (PTE). The seven horizontal “periods” reflect the main electron
shells, with n = 1–7. The 18 vertical columns form “groups” of elements with similar chemical
properties (because of similar outer electron shell occupancies). Groups 1, 2, and 13–18 are the
“main” groups.16 The value below each element’s name is its molecular weight in g/mol.
If the excited electron reaches a sufficiently high energy level to escape from the
nucleus’ attraction, this energy value is the ionization energy Ei(e). In the case of the
hydrogen atom, this value is 13.6 eV. (see Tab. 1.3).
n ... ∞
Color-coded substructures of periods mirror the subshells s, p, d and f, with maximum occupan-
cies of 2, 6, 10 and 14 electrons, respectively. This can be explained by quantum number systemat-
ics, see below.
1.4 Quantum mechanics 17
Within the BOHR atom model, not only does each electron have a characteristic en-
ergy as defined by its shell, but each shell was also postulated to accept a specific
number of electrons. Moreover, although multiple electrons can occupy the same
main shell, each electron of an atom is still considered to be unique. Of the 76 elec-
trons located in various shells of the gold atom, all are different in at least one
parameter!
18 1 The atom’s structure I: Electrons and shells
2400
He
2200
Ne
2000
1800
Average Ionization Energy (kJ/mol)
1600 Ar
1400 Kr
H Cl Xe
1200
P Rn
1000
S
800 Si
Mg
600
Al
400 Li Na
K Rb
Cs
200
0
10 20 30 40 50 60 70 80 90
Atomic Number
Fig. 1.11: First ionization energy of the first six periods of chemical elements. Within a group (e.g.
the noble gases, red line), ionization energies decrease with increasing proton number: He (Z = 2)
! Ne (Z = 10) ! Ar (Z = 18) ! Kr (Z = 36) ! Xe (Z = 54) ! Rn (Z = 86). Within a period, ionization
energies increase; e.g. for the n = 3 (L) period (blue line): Na (Z = 11) ! Mg (Z = 12) ! Al (Z = 13) !
Si (Z = 14) ! P (Z = 15) ! S (Z = 16) ! Cl (Z = 17) ! Ar (Z = 18).
The nomenclature applied is “quantum numbers”. It bridges the energy levels of the
BOHR atom model with quantum mechanics introduced by SCHRÖDINGER et al. and molec-
ular orbital theory proposed by HUND and MULLIKEN. Accordingly, there is a set of four
quantum numbers which characterize each electron, like a fingerprint. Three (n, l, m)
are dedicated to the orbital parameters of the electron’s existence in terms of energy and
orbital angular momentum, and the fourth one (s) to the intrinsic spin of the electron.
The principal (or main) quantum number is n. It defines the different main
shell the electrons exist. It quantifies the maximum number of electrons to fit into
each main shell. The energy of these shells, however, depends on the individual
atom. There are basically seven principal quantum numbers with n = 1, 2, 3, 4, 5, 6,
7, also symbolized by Latin capital letters K, L, M, N, O, P, Q. Artificial elements of
Z > 118 will start filling electrons in the n = 8 shell.
Shells may be structured into subshells. All shells except the K-shell (n = 1)
involve multiple subshells. The subshells are indicated by the orbital quantum
number l.
1.4 Quantum mechanics 19
The value of the electron’s orbital angular momentum reflects individual orbits
of electrons within a given main shell. Thus, many electrons may “circulate” within
one and the same main shell, but do not overlap, as they populate individual or-
bital spaces.
The orbital angular momentum correlates with the principal quantum number.
The higher the principal quantum number n, the more subshells may be populated
within that shell. The maximum number of subshells is l = n − 1, i.e. 0 ≤ l ≤ (n − 1).
Consequently, there is l = 0 for n = 1 (because n − 1 = 0). In contrast, there are four
subshells for n = 4, namely l = 0, l = 1, l = 2, l = 3. Those orbital orbitals are denoted
as s, p, d, and f.17
Quantum mechanics not only introduced specific energetically “allowed” orbi-
tals for the circulating electrons, it also characterized electrons in more detail. The
key point was to consider electrons not as localizable particles like planets, which
can be correlated at any point in time to a specific position. Instead, their existence
in space and time was proposed to follow probabilities only, defining states of high
and low probability. Because only discrete values may exist in quantum mechanics
(as indicated by the PLANCK constant,18,19), the “real” orbital angular momentum L
is identified through the relation
2
h
L2 = (1:4)
lðl + 1Þ
h
h = (1:4a)
2π
Figure 1.12 illustrates the different features of electrons with different principal and
angular quantum numbers existing in space along an x-y coordinate system.20
All s-electrons (l = 0 orbitals) exist in sphere-shaped orbitals around the nu-
cleus. The distribution within that sphere is homogeneous, i.e. there is a certain
probability for the s-electron to be in the center of the coordinate system.21 Because
This quantum physical idea can predict “real chemistry”, i.e. the chemical behavior of ele-
ments. The PTE shows a specific block of “p-elements” containing elements with the outer electrons
located in the p-shell of the main shells n > 1; these are the main PTE groups III through VIII. There
are also “d-elements” and “f-elements”, i.e. the lanthanides and actinides. Again, experimental evi-
dence is beautifully correlated with theoretical concepts on the atom’s electronic structure.
The symbol ħ stands for the “reduced” PLANCK constant, where ħ = h/2π. The reduced PLANCK
constant is preferred in cases where radial dependencies are relevant, such as angular frequency,
solid angle, etc. (see Fig. 1.20).
Nobel Prize in Physics to PLANCK, 1918, “in recognition of the services he rendered to the ad-
vancement of Physics by his discovery of energy quanta”.
This is mathematically derived for the simplest case – the one electron system of hydrogen.
One of the main features to understand nuclear transformations is the principal difference be-
tween probabilities of existence of electrons of s orbital quantum number compared to other orbital
quantum numbers. As depicted in Fig. 1.12, s orbital electrons do have a probability (although
20 1 The atom’s structure I: Electrons and shells
3s 3p 3d
2s 2p
1s
Fig. 1.12: Two-dimensional illustration of the orbital distribution of s, p, and d electrons for n = 1–3,
imaged as densities around the atom’s nucleus. For l = 1 (1s, 2s, 3s, etc. electrons), the density
distributions are spherical. The sphere’s radius increases slightly with n, following the relationship
in Fig. 1.7 and according to eq. (1.1). The change in size is more dramatic as Z increases, when
electrons become more and more attracted to the nucleus. For instance, the 1s-electron gets closer
to the nucleus if the chemical element has a larger proton number and electrons are filled in 2s
(i.e. starting from lithium) or 3s (i.e. starting from sodium) orbitals. The same is true for p-orbitals
of n = 2 and n = 3 shells (2p, 3p), etc.
the electron binding energy differs between shells (see Tab. 1.3), the radial distance
between each orbital and the center of the coordinate system increases with n. For
example, the s-electrons of the n = 2 shell (L) are further from the nucleus than n = 1
(K) electrons.
extremely small) to exist close to the center of the Cartesian coordinate system (i.e. the nucleus);
shells with different orbital quantum numbers do not. Despite the extremely large dimensions of
electron shells compared to the atomic nucleus, s shell electrons are “allowed” to exist close to and
even within the atomic nucleus. Of course, these probabilities drop dramatically when increasing
the principal quantum number from n = 1 to n > 1. At minimum, each atom will have K-shell s-
electrons, and important nuclear transformations of radioactive nuclides (such as the electron cap-
ture branch of the β process or secondary processes like inner conversion) rely on this feature of s
orbital electrons (see Chapters 8 and 11).
1.4 Quantum mechanics 21
From l > 0 onward, the electron density distribution is no longer spherical. The
orbital distribution of p-electrons (l = 1) is dumbbell shaped and the d-electrons (l = 2)
exist in orbitals of other shapes (see Fig. 1.12). Thus, a three-dimensional system is
needed. Figure 1.13 illustrates the different electron features of p-orbital electrons ex-
isting in space along an x-y-z coordinate system.22 The general profile is dumbbell
shaped. The two symmetrical orbitals on the left and the right represent so-called
bonding or nonbonding parts. Within a Cartesian system, there are three different op-
tions for a dumbbell: a p-electron of identical characteristics may either occupy space
along the x-axis, the y-axis, or the z-axis (px, py, pz). In the center of the coordinate
system, i.e. at x = y = z = 0, the probability to exist is zero.
z z z
x x x
y px py pz
y
Fig. 1.13: The three-dimensional dumbbell-shaped orbitals of p-electrons along x-y-z axes of a
Cartesian coordinate system.23
For l = 0, there was one s orbital. For l = 1, there was one p, but with three individual
profiles: px, py, pz. The number of those subshells increases with increasing values of
l. Mathematically, it is expressed by − l, −(l − 1), . . ., 0, . . ., +(l − 1), + l (see Tab. 1.4).
For l = 2, for example, there are 5 possible states, namely − 2, − 1, 0, + 1 and + 2.
Each given l-orbital thus contains a set of electrons described by a new parame-
ter: their magnetic quantum number m. It may be thought of as if electrons within a
certain subshell may form groups of “clouds”. In fact, it reflects the projection of
the orbital angular momentum of the electron within a magnetic field, referred to as
an axis of the polar coordination system. (Negatively charged electrons moving
around a positively charged center are influenced by an electromagnetic field.) The
final orbital angular momentum thus becomes Lz = ħml.
This again is mathematically derived for the simplest case – the one-electron system of
hydrogen.
This organization of orbital distribution of the 3p-electron states is not an abstract concept. It is
reflected in real chemistry, where the px, py, pz states make a huge difference in terms of forming
chemical bonds and thus forming chemical compounds.
22 1 The atom’s structure I: Electrons and shells
Tab. 1.4: The four quantum numbers making each electron of an atom unique.
Orbital angular l ≤ l ≤ (n − ) , . . ., n − s, p, d, . . . l = , , . . .
EARTH
Earth spinning
SUN around its axis
(spin angular momentum)
Fig. 1.14: Conception of the two different angular moments of the planet earth circulating around
the sun. The earth’s movement in space, taking one year, would mirror the orbital angular
momentum of an electron around an atom’s nucleus – reflected by the quantum number
l. Similarly, the earth’s own movement around its axis, lasting one day, may help to understand the
spin angular momentum of the electron – denoted by the spin quantum number s.
However, the momentum is either in one direction or the other – there are no further
options. The corresponding spin quantum number s thus is either +½ or −½.24 It re-
flects an intrinsic parameter of the electron. For example, for a given set of n, l, m
Note that there is actually no “positive” or “negative” direction of the angular spin.
1.4 Quantum mechanics 23
there are always two options for an electron, which differ in angular spin of +½ or
−½. This is denoted as HUND‘S rule. The value of this spin quantum number s is:
Now each electron is fully identified by (n, l, m, +½) and (n, m, l, −½). Each subor-
bital, defined by (n, l, m), can contain a maximum of two electrons of opposite spin,
and the general notation for how magnetic quantum number and spin are linked
is ms = ±½. Finally, each electron is supposed to be unique. This is known as the
PAULI exclusion principle:25 in any quantum mechanical system (such as electrons
in the atom’s shell, nucleons in the atom’s nucleus), two particles of identical quan-
tum parameters should not exist.
Figure 1.15 illustrates how electrons fill the individual orbitals for the element cal-
cium. The complete composition of this element of Z = 20 according to quantum
number logic would 1s2 < 2s2 < 2p6 < 3s2 < 3p6 < 3d2. However, this order is followed
only for the first 18 electrons. The two outer electrons, i.e. electron numbers 19 and
20, prefer to occupy the 4s2 orbital instead of the 3d orbital! The explanation lies in
the energy of the individual shells and subshells. It is a matter of fact that the 4s2
“comes first”, i.e. shows a lower value of EB(e) relative to the binding energy of the
The Nobel Prize in Physics, 1945, was awarded to W PAULI “for the discovery of the Exclusion
Principle, also called the Pauli Principle”.
24 1 The atom’s structure I: Electrons and shells
4d 4d 4d 4d 4d empty
shells
3s
2s
Energy
1s
Fig. 1.15: Electron shells arranged according to the system of the four quantum numbers are filled
by electrons according to the electron binding energy of each subshell. (left) Electron orbitals
defined by quantum numbers and arranged with increasing electron binding energy. (right) The 20
electrons of the chemical element calcium filled into main shells (circles), subshells not indicated.
The chemistry of calcium is defined by its outer electrons, which are the two 4s electrons instead
of 3d electrons.
3d2 subshell electrons.26 This effect becomes more pronounced the closer the outer
electron shells are in electron binding energy. Furthermore, because differences be-
tween individual subshells of higher main quantum number become smaller and
smaller (see Fig. 1.7), the effect of overlapping subshell orbitals becomes more and
more pronounced.
This order of electron subshell occupancies is reflected in the structure of the PTE, i.e. it per-
fectly correlates the experimentally observed systematics of chemical elements with electron struc-
ture parameters. Interestingly, the same will be discussed in the following chapters, where the
overlap of nucleon shell structures will explain the stability of the nuclei of atoms.
1.5 Mathematical explanations of the BOHR atom model 25
BOHR’S atom model postulates that electrons are located in various shells. This
statement satisfies experimental evidence, such as line spectrum of hydrogen and
positioning of chemical elements within the PTE. However, mathematical verifica-
tion of the model’s postulate was delivered only later. SCHRÖDINGER27 developed the
corresponding mathematics to calculate parameters such as the electron energy of
the simplest model, the one electron of atomic hydrogen. The concept is to quantify
electron parameters in terms of particles existing within a potential well.28
The key intention is to deduce fundamental parameters of the electron (and
other particles), such as energy (potential and kinetic), velocity, and position. The
mathematics itself in detail is beyond the scope of this book. The basic idea is to
treat the electron as a wave, and to derive its wavefunction Ψ. Finally, the quantum
mechanical parameters are correlated to the basic physics of the electron, such as
the radius r of a certain electron shell, and its energy described as E = f(r).
A potential well considers a particle (here, an electron29) inside a well (or box), and
defines a region defined by a (local) minimum of potential energy. In nature, a lake
filling a valley surrounded by mountains represents a three-dimensional potential
well.30 For mathematical treatments, one-dimensional model systems are consid-
ered. The well localizes a “free” particle within the two walls of (infinite) high po-
tential energy along an x-axis. Inside the box, the potential energy U is zero, while
outside it is infinitely large.31 The two walls are characterized by positions 0 and L
Nobel Prize in Physics, 1933, “for the discovery of new productive forms of atomic theory”.
It is introduced here because it is also relevant to the quantification of nucleons within a poten-
tial well. It is thus applicable to particles of the atom’s nucleus, for e.g. understanding the nucleon
shell occupancies (see Chapter 2), and introducing the tunneling effect (see Chapter 9).
. . . could also be a gas molecule, an electron, or a nucleon.
The simplest model implies a single, negatively charged “free” electron existing in a box (“par-
ticle in a box”) within the electrostatic attraction caused by the positively charged nucleus with a
potential energy V > 0. This is a potential well, which does not, however, completely describe the
whole truth of electron(s) within a real atom.
There are slightly modified considerations for a potential well.
26 1 The atom’s structure I: Electrons and shells
(see Fig. 1.16). Within the well, i.e. between positions 0 and L, the particle moves
with constant velocity and is reflected from each wall without losing energy. It can
leave (or escape the well) only if it becomes energetically excited.
U=
U=
U (x)
O L
x
Fig. 1.16: Electron potential well. An electron inside a well (a box) exists within a specific region
defined by a (local) minimum of potential energy. The well localizes the “free” electron within two
walls of (infinitely) high potential energy along an x-axis of distance. Inside the well, the potential
energy U is zero, while outside it is infinitely large. The electron moves with constant velocity
between the two borders (walls), characterized by positions 0 and L.
n2 h 2 n2 h2
En = · = · , with n = 1, 2, 3, . . . (1:6)
2L2 m 2L2 4m
πn
ν= (1:7)
L
1.5 Mathematical explanations of the BOHR atom model 27
n=1 =½
n=2 =1
n=3 = 3/2
Fig. 1.17: Waves being reflected at the walls of a potential well. A wave may propagate “forever”
only if half its wavelength λ is an integer multiple n of the well length L: Lwell = f(n ½λ).
exact position
impossible to
determine
position
potential
probability of
Fig. 1.18: The harmonic oscillator in classical physics and in quantum mechanics.
^ = EΨ
HΨ ^ (1:10)
^ = − 1 h ∇2 + U
H (1:11)
2m
^ = ih δ
E (1:12)
δt
1 h 2
− ∇ Ψ ðx, y, zÞ + Uðx, y, zÞΨ ðx, y, zÞ = EΨ ðx, y, zÞ (1:13a)
2m
The HAMILTONIAN operator in its general form is expressed in eq. (1.11), and com-
bines terms for both kinetic energy ½ħ∇2 and potential energy U. For the kinetic
energy, the reduced PLANCK constant ħ and the mass of the particle are considered.
Further, the nabla operator ∇ is involved,33 which represents the derivatives for all
coordinates in space. Time is not considered here.
The energy operator also contains ħ. The imaginary unit i represents the square
root of − 1, which breaks wavefunctions into a real component and a virtual compo-
nent. Ê includes a derivative for time only, and does not consider position (x, y, z).
Solutions of the energy operator represent sine functions of characteristic frequencies.
Solutions to eq. (1.13a) are wavefunctions, Ψ, that satisfy the parameters of
both operators. In the case of complex wavefunctions of type Ψ(r, t), the harmonic
potential correlates with the eigenfrequency ω of the harmonic oscillator. The corre-
sponding HAMILTONIAN function H depends on the mass m and the impulse p and
reflects the total energy of the system composed of kinetic and potential energy.
The time-independent (i.e., position-related) SCHRÖDINGER equation is given in
eq. (1.13b). Here, the three coordinates (x, y, z) are represented by the radius r. In
the specific case of the x-dimension, solutions to the eigenvalue equation for the
stationary SCHRÖDINGER equation are the eigenvalues Én as in eq. (1.14). In this case,
Δ is the LAPLACE operator, where Δ = δ/δx in a one-dimensional case.
In the case of a one-electron system, this equation finally yields a description of
the electron’s orbitals.
The SCHRÖDINGER equation yields the probability of finding (the residence) of a par-
ticle, which is expressed as Ψ 2. For the position of the electron within the harmonic
oscillator, there is one probability only, which is the square of the wavefunction, i.e.
Ψn(x)2. Figure 1.19 reproduces wavefunctions of the electron in a harmonic oscillator
for n = 1–7 and the corresponding probabilities of presence or “probability density”.
The orbital energy En is found using En = ħω. The harmonic oscillator accepts only
discrete energies where n is a whole number. The lowest possible state is Eo = ½ħω,
where Eo > 0. This is known as the zero-point energy, or the energy of a ground-
state electron.
1 h 2
− Δ Ψ ðrÞ + UðrÞΨ ðrÞ = EΨ ðrÞ (1:13b)
2m
1 2 1
É n Ψ n ðxÞ = − h ΔΨ n ðxÞ + mω 2 xΨ n ðxÞ (1:14)
2 2
1
En = hω n + (1:15)
2
Unlike classical mechanics, which provides a particle’s exact position and exact im-
pulse p, in quantum mechanics the HEISENBERG uncertainty principle applies. This
principle states that an electron’s exact position x and exact impulse p cannot be
simultaneously known: ΔxΔp ≥ ħ/2.
30 1 The atom’s structure I: Electrons and shells
Ψ₇(x)
Ψ₆(x)
Ψ₅(x)
Ψ₄(x)
Ψ₃(x)
Ψ₂(x)
Ψ₁(x)
Ψ₀(x)
(a)
Ψn(x)²
(b)
Fig. 1.19: Wavefunctions of a quantum harmonic oscillator (a) and probabilities of presence Ψn(x)2 (b).
h
Δ xΔ p ≥ (1:16)
2
1 ðΔpÞ2 1
En = + mω2 ðΔxÞ (1:17)
2 m 2
1 ðΔpÞ2 1 mh2 ω2
En ≥ + (1:18)
2 m 8 ðΔpÞ2
1.5 Mathematical explanations of the BOHR atom model 31
The harmonic oscillator models a particle in one dimension. In reality, the hydrogen
electron surrounding its proton is three-dimensional. In this case, the SCHRÖDINGER
equation uses a polar coordinate system instead of Cartesian coordinates. Any point
in space is defined as the distance r (a vector) of a point from the center of the sphere,
as well as two angles φ, θ between the vector and the directions of the Cartesian co-
ordinates. These angles define a so-called solid angle Ω (see Fig. 1.20).
dφ
φ
dγ
dΩ
Fig. 1.20: The solid angle Ω describes a two-dimensional angle in three-dimensional space. Along
with the distance vector r, it is part of a polar coordinate system.
These three polar variables (r, φ, θ) are treated separately by wavefunctions, which
are divided into one part for radial distance Rn,l and another one for angular posi-
tions Yl,m. These are specific to one electron as characterized by its three quantum
numbers n, l, and m.
Finally, a BOHR shell electron may be described with a set of “real” physical param-
eters, such as position, impulse, and energy. The radius r gives the radial distance
between the electron and nucleus. The potential energy U of an electron thus de-
pends on r. Because this radius depends on nuclear charge, another parameter
needed is Z: the number of protons in the nucleus. The potential energy U(r) follows
a ratio between Z and r, and is given in eq. (1.20) proportional to the electric charge
density eρ.34 Equation (1.21) then combines the kinetic energy ½ħ(∇)2 and potential
energy U in a simplified version following eq. (1.13b).
Ze2ρ
UðrÞ = − (1:20)
n
eρ = the electrical charge density. The charge of an electron is −1.602⋅10−19 C. The charge den-
sity is referred to as the volume of the electron, with units in C/m3.
32 1 The atom’s structure I: Electrons and shells
Zeπ 2
Ekin − Ψ = EΨ (1:21)
n
Expected eigenenergy values of the electron located within one of the main shells,
denoted by E, are different between spherical orbitals (l = 0, i.e. s-electrons) and
cases where l > 0. This equation finally corresponds to eq. (1.1) and provides the math-
ematical background of important electron parameters of the BOHR atom model.
n2 h2
Enl = En0 = (1:22)
8mL2
1.6 Outlook
The electron structure of the BOHR atom was deduced from the hydrogen line spec-
trum, and the SCHRÖDINGER equation calculates the parameters of the one-electron
system. With an increasing number of electrons, however, the interactions between
electrons and between electron and nucleus should be considered. Even for the
one-electron system, there are a number of effects originating from a more sophisti-
cated structuring of electrons inside the shells, with some of them being responsible
for fundamental physico-chemical effects and analytical techniques. These are sub-
sumed under “fine structures”. Fine structures, more generally, are created by spin-
orbit interactions and reflect the difference between nonrelativistic and relativistic
treatments of the electron energies. There is a correction term for relativistic effects
of the kinetic energy of the electron (to be considered in the case of objects moving
with velocities close to the speed of light), and there is one for the potential energy
of the electron (DARWIN term). This turns even the hydrogen line spectrum into a
rather difficult, complex system.
Dimensions of fine structures (emission line splitting) follow a fine-structure con-
stant α (where α = 7.297⋅10−3) and correlate with the proton number Z through Z2α2.
In the case of hydrogen, the effects are very minor and could experimentally be re-
solved relatively late.35 Fine structures of the α-line of the BALMER series (656.28 nm
wavelength) amount to only 0.014 nm (in dimensions: 0.14 Å vs. 6562.8 Å).
WE LAMB: Nobel Prize in Physics, 1955, “for his discoveries concerning the fine structure of the
hydrogen spectrum”.
1.6 Outlook 33
In the context of this textbook, the spin-orbit coupling, which contributes to elec-
tron line fine structures, is of special interest.
Parameters like orbital angular momentum and spin angular momentum can be
attributed to electrons. The two individual angular momenta may overlap within an
electromagnetic field. This is called spin-orbit interaction or spin-orbit coupling. Or-
bital angular quantum l and spin angular quantum momentum s numbers create a
total angular momentum quantum number j with j = |l + s|. The individual spin quan-
tum number remains ±½, but becomes arranged as illustrated by a rotating vector
cone (see Fig. 1.21). The total angular momentum J is related to j through eq. (1.23).
2ħ
ħ
J=L+S
-ħ
-2ħ
Fig. 1.21: Spin-orbit coupling. Total orbital angular (L) and total angular spin number (S) moments
create a total angular momentum number J.
To demonstrate the calculation of j, consider the following example. For orbital quan-
tum numbers l = 1 containing six p-electrons (assuming the three states of magnetic
quantum numbers m = 1, 0, − 1, (i.e. px, py, pz) are each filled by two electrons of spin
quantum number + 1/2 and − 1/2, respectively), there are two possible values of j: j =
1 + 1/2 = 3/2 and j = 1 − 1/2 = 1/2. The electrons will reach two different levels, namely
j = 2 + 1/2 = 5/2 and j = 2 − 1/2 = 3/2. There will always be two values of j, independent
of the specific orbital quantum number considered, except when l = 0.
The number of electrons that fit into the two new suborbitals is 2j + 1. For exam-
ple, a fully filled p-orbit distributes its altogether 6 electrons in a ratio of 4:2,
34 1 The atom’s structure I: Electrons and shells
ð2l + 1Þ
Δ Els = Uls ðrÞ (1:24)
2Vls ðrÞ
Again, this theoretical concept was proven by experimental data. Emission line
splitting following Z2α2 is, for instance, “obvious” in the case of emissions of so-
dium lamps (Z = 21). Sodium lamps for street lighting show a typically yellowish
bright light. Let us take a deeper look into that emission.
The electrons of the sodium atom can be excited following the simplest model
for the hydrogen line spectrum. The valence electron of sodium, in a 3s1 configuration,
may reach the “free” 3p subshell level through excitation. From that level, it returns
to the 3s orbital by emitting the electron binding energy difference (between the 3s
and 3p levels) as a photon. Because of spin-orbit coupling within the 3p shell, how-
ever, the 3s electron (instead of just one 3p level) traverses two specific p-electron en-
ergy levels, represented by the two total angular momentum numbers 3p1/2 and 3p3/2.
As a result, sodium emits two discrete photons of wavelength D2 = 588.9950 nm
and D1 = 589.5924 nm. Although the difference in binding energy between the two
degenerate 3p levels is only ΔE = 0.0021 eV, it is clearly a line doublet (Fig. 1.22).36
1.6.3 Multiplicity
Each individual s, p, d, and f suborbital may accept two electrons. The electron or-
bitals are filled by electrons according to the HUND rule. In the case where several
degenerate levels (wavefunctions) are available, such as the three px, py, and pz
subshells, electrons are filled in a way to maximize overall spin values.
The different versions of occupancy of degenerate subshells are called multi-
plicity M. This quantum mechanical unit thus refers to the possible number of ori-
entations of the angular spin:
M = ð2S + 1Þ (1:25)
This experimental observation for the Na D-line and of similar effects in the case of other alkali
(s1 orbital) elements actually motivated SA GOUDSMIT and GE UHLENBECK in 1925 to postulate that
electrons have an intrinsic spin angular momentum.
1.6 Outlook 35
E = 0.0021 eV
p½ p½
3p p p p j(j+1)
589.0 nm
589.9 nm
3s s s
Fig. 1.22: Fine structure within the 3p level of sodium, caused by spin-orbit coupling, in terms of
energy (a) and capacity (b).
Paired electrons, e.g. two electrons of spin +½ and −½ located within one filled or-
bital, compensate for their interaction. Consequently, only the number of unpaired
electrons and their individual angular spins matter. According to eq. (1.25) and
knowing S must be a multiple of ½, multiplicities of 1, 2, 3 . . . (called singlet, dou-
blet, and triplet states) are possible. Systems with no unpaired electrons always
exist in a singlet state, those with one unpaired electron in a doublet state, and
those with two exist in a triplet state. Higher multiplicities are also possible. Multi-
plicity notation is shown in Fig. 1.23.
MULTIPLICITY
NUMBER
OF ELECTRONS
SUB-SHELL (l)
Fig. 1.23: Notation of multiplicity M. The multiplicity is indicated as left superscript for each
subshell of quantum number l (in the case: l = 1 = p). The number of electrons involved is the right
superscript.
Let’s consider another example. The carbon electron shell (1s2 + 2s2 + 2p2) is filled
pairwise for the 1s and the 2s shells, but the 2p shell is filled according to 2(px)1 + 2
(py)1 instead of 2(px)2. In the latter (hypothetical) case with all electrons paired, M
would be (2 × 0) + 1 = 1, indicating a singlet state. However, the real situation gives
M = (2 × ½) + 1 = 2, indicating a triplet state.
36 1 The atom’s structure I: Electrons and shells
Aim: While the first chapter recalled the principal composition of an atom, consisting
of a nucleus and a shell with electrons of individual properties, this chapter introduces
the principles and components of the atom’s nucleus structure. The nucleus contains
protons and neutrons, collectively called nucleons. Nuclei with the same number of
protons but different numbers of neutrons create a rich world of isotopes of the same
element. Every nucleus is described by a set of three numbers: the number of protons
(identifying the chemical element), the number of neutrons (reflecting the specific iso-
tope of that element), and the sum of protons and neutrons (forming the mass num-
ber). Analogous to the Periodic Table of Elements, the Chart of Nuclides arranges
chemical properties of elements according to their electron shell parameters along
with increasing mass of the element. The Chart of Nuclides correlates more than 3000
nuclei identified so far in an (x,y)-coordinate system, with x = neutron number, N and
y = proton number, Z.
Mass and volume (or radius) are the physical properties of every set of nucle-
ons. Mass divided by volume gives the density of a nucleus – the density of matter.
Radius may refer to the (homogeneous) mass distribution within the nucleus’ vol-
ume, but also to charge distribution (which is the distribution of protons within the
nucleus).
The mass of individual nuclei is of ultimate interest since mass is proportional
to energy. Exact knowledge of the precise mass of a nucleus provides access to the
overall nucleon binding energy. The binding energy per nucleon determines, to a
significant extent, the stability or instability of a given nucleus.
With quantum mechanics, the one electron of the hydrogen atom was finally under-
stood as energetically “surviving” in specific orbits of the shell. The electron is at-
tracted by the one proton in the nucleus of the hydrogen atom. This particular atom
has no other components. However, the atoms of all other chemical elements con-
tain an additional basic constituent – the neutron.
In contrast to the electron and the proton, which were identified according to
their charge and mass by the deviations they experience in magnetic fields, the neu-
tron was identified much later, only in 1932. J CHADWICK1 positioned an α-radium
source close to a beryllium sample (ZBe = 4), as RUTHERFORD had done with gold foil
(ZAu = 79). But unlike the RUTHERFORD experiment, the α-particles interacted with the
https://doi.org/10.1515/9783110742725-002
38 2 The atom’s structure II: Nucleons and nucleus
beryllium nucleus through a nuclear reaction, forming a (new) carbon nucleus and an
unknown particle – the free neutron.
Both products were not easy to analyze: the newly formed atom, carbon, was
uncharged, stable, did not emit any radiation, and the neutron released was also
electrically neutral. The neutron was detected rather indirectly, namely in terms of
a follow-up nuclear reaction between the neutron and a stable isotope, such as hy-
drogen or boron. Boron was an ideal element for this purpose because of its very
high probability to catch a neutron (see Fig. 2.1).2 The energy and mass of the neu-
tron could be derived e.g. from the transfer of the impulse of the neutron to the
proton.
PROTON
DETECTION
Fig. 2.1: Nuclear processes leading to the discovery of the neutron. The new radiation emitted from
this process was historically called “beryllium radiation”.
See Chapter 13 for nuclear reaction mechanisms and cross-sections of neutron capture.
Interestingly, proton and neutron mass differ by a small fraction – the neutron is about 0.1%
heavier than the proton. This seemingly negligible mass excess of the neutron, however, becomes
the reason for a fundamental difference between the two particles. The “free” proton, i.e. one that
is not bound in an atomic nucleus, is stable, while the analogue “free” neutron is not. Instead, it spon-
taneously transforms into a proton (and other products). For details, see Chapters 4 and 8.
2.2 Nuclide notations and isotopes 39
smaller by a factor of ca. two thousand. The overall mass of an atom is thus pre-
dominantly localized in its nucleus.4 On absolute scale, the mass of the three par-
ticles is extremely small. About 6 × 1023 protons or neutrons would weigh only one
gram.
The classical properties of these particles, i.e. mass, charge, charge-to-mass ra-
tios, but also relative mass in terms of atomic mass unit (see below) and mass ex-
pressed in terms of energy are summarized in Tab. 2.1. According to the equivalency
of mass and energy, the energy is 938.272 and 939.566 MeV for the proton and the
neutron, respectively.
Tab. 2.1: Summary of basic properties of the three basic atomic constituents: the electron, proton,
and neutron. The elementary charges are − 1, + 1, and 0 for the three particles, but are given as
values of the elementary (positive) charge unit C (coulomb). For the electron and the proton, the
ratio between charge (q in C) and mass (m in g) was experimentally derived. The mass is given in
units of kilograms and u (the atomic mass unit), as well as in energy in MeV via E = mc2. (For
energies relevant to nuclear physics, the unit MeV is preferred to the Joule J.5)
C/g C kg u MeV
Electron −.⋅ − −.⋅− .⋅− . .
Proton + .⋅ + +.⋅− .⋅− . .
Neutron .⋅− . .
With the two constituents of an atomic nucleus, the proton and the neutron, it is
now possible to describe an atom not only by its “atomic number” (which is the
proton number, Z) and atomic mass (denoted in the PTE as g/mol) but also by its
nuclear composition. The standard notation uses the element’s symbol (E), adding
the number of protons (Z) and neutrons (N) as subscripts (on the left- and right-
The mass number of natural gold is 197 and the proton number is 79. A simple estimate
would say that there are 79 protons + (197 − 79 = ) 118 neutrons in the nucleus, making a mass of
79 × 1.673⋅10 −27 kg + 118 × 1.675⋅10 −27 kg = (132.167 + 197.650)⋅10 −27 kg = 329.817⋅10 −27 kg for
the nucleus versus 79 times the mass of an electron = 0.07⋅10−27 kg. With an overall mass of
329.887⋅10 −27 kg, the nucleus of a gold atom represents about 99.98% of the entire atom’s
mass. (This simple mathematics is not completely true, as the mass of “bound” nucleons differs
from the sum of individual, unbound nucleons; see later in this chapter.)
See Appendix for all the parameters listed in the book.
40 2 The atom’s structure II: Nucleons and nucleus
Mass
Mass number
number A
A
A Element symbol E
Fig. 2.2: Notation showing the numbers of protons (Z), neutrons (N), and nuclear mass (A) of a
chemical element (E). The nucleus of the helium-4 atom, i.e. the α-particle, is shown as an
example.
In practice, it is sufficient to denote an isotope using only the mass number, be-
cause the chemical element and proton number are inherently linked. Thus, it is
straightforward to calculate the corresponding neutron number N of the given iso-
tope. For example, the α-particle, representing the nucleus of the helium atom and
containing two neutrons, may simply be written as 4He. It is composed of two pro-
tons plus two neutrons (Z = 2, N = 2, and A = 2 + 2 = 4). The α-particle thus represents
the helium atom without its two electrons.6
Note that Z, N, and A do not represent absolute mass numbers. The real mass
would result from multiplying the individual proton and neutron numbers with the
absolute mass of each nucleon, as given in Tab. 2.1.
2.2.2 Isotopes
For the hydrogen atom discussed, the nucleus contained only one proton (and no
neutrons). With the atom ionized, this nucleus is simply a “naked” proton (see
Fig. 2.3). However, there exists another stable hydrogen nucleus, having one proton
It is created by ionizing the atom, i.e. by transferring enough energy to the helium atom to re-
lease both of its 1s electrons above the ionization potential. Similarly, the “naked” nucleus of
atomic hydrogen is the proton itself (see Fig. 2.3).
2.2 Nuclide notations and isotopes 41
The proton
1 2 3
1.00794
9
Fig. 2.3: Isotopes of hydrogen, expressed using the number of protons, neutrons, and mass for
each of three isotopes. The hydrogen isotopes containing one or two neutrons are deuterium and
tritium. The average molecular mass number of naturally occurring hydrogen is 1.00794 g/mol.
and one neutron. Helium also has two stable nuclei, having either two neutrons
(the α-particle) or one neutron.
Thus, Z remains the same (Z = 1 for H, Z = 2 for He) since it characterizes the ele-
ment, and (due to the electron configuration defined by Z) its chemical reactivity is
also unchanged. However, these hydrogen and helium nuclei differ in their respec-
tive number of neutrons (by N = 0 and 1 for H, and N = 1 and 2 for He). Nuclei of
constant Z but varying N are called isotopes, as they are “iso” (equal) in Z number.
Consequently, isotopes differ by mass (as A = Z + N) according to A = 1 and 2 for H
and A = 3 and 4 for He. In general, isotopes are written with the element symbol
only (see Fig. 2.3).7
Because of the relative difference in mass (deuterium is roughly twice as heavy as
hydrogen), the three hydrogen isotopes (and their chemical species) differ profoundly
in some physico-chemical properties. Within the same element, these differences be-
come smaller with increasing Z, due to the decreasing differences between the rela-
tive mass of the isotopes.8
In the case of hydrogen, unlike all the other chemical elements, the other stable (2H) and the
main radioactive (3H) isotopes were given individual names and symbols: deuterium D (due to the
mass number 2) and tritium T (due to mass number 3).
Still, there are detectable effects even for the heavy chemical element uranium. For example, the
small differences in mass in uranium hexafluoride gas UF6 are used to separate different isotopes
42 2 The atom’s structure II: Nucleons and nucleus
The two stable hydrogen isotopes, H and D, were formed at the element genesis
in vastly different ratios. The third isotope, radioactive T, is permanently being pro-
duced in the earth’s atmosphere by nuclear processes induced by cosmic radiation.
The mixture of the three hydrogen isotopes present on our earth follows the percent
distribution of 99.9885%, 0.0115%, and about 10−15% for H, D, and T, respectively.
The overall mixture of the three hydrogen isotopes is dominated by the lightest iso-
tope (H), with small contributions from D and a practically negligible contribution
from T. The resulting overall mass number is 1 × 0.999885 + 2 × 0.000115 × 2 + 3 ×
10−17 = 1.000115, i.e. a bit higher than the most abundant isotope H. The resulting
average molecular weight is 1.00794 g/mol.
Mixtures of naturally occurring isotopes of stable elements thus generate average
mass numbers. In other words: an element’s mass number reflects the abundance of its
constituent isotopes. Similarly, the element’s molecular weight represents an average
value. However, a few chemical elements consist exclusively of one stable isotope.9
2.3.1 Radius
After having described the three principle subatomic particles, the parameters of the
atom can now be discussed in more detail. In general, there is a “spherical” model of
the atom (see Fig. 2.4 for the main isotope of carbon, 12C). The dimension of the
(outer) electron shell generally defines the radius r of an atom.10 Despite a few excep-
tions, there are systematic trends regarding radius size within the PTE. In the same
group of the PTE, radius increases from top to bottom. Within the same period, the
radius decreases from left to right. The latter is due to the increasing proton number
Z within the same principal electron shell (i.e. the main quantum number n) and is
called contraction. Hydrogen is thus understood to be the smallest atom, while the
heaviest alkaline element (francium) is understood to be the largest. For hydrogen, r is
around 32 pm; for a heavier element such as gold it is about 144 pm, and for cesium –
the heaviest stable alkaline element and the lighter homologue of francium – it is
of uranium (e.g. the isotopes 235 and 238 needed for nuclear fission) out of the isotopic mixture of
natural uranium by gas centrifuge cascades.
Beryllium (9Be), fluorine (19F), sodium (23Na), aluminum (27Al), phosphorus (31P), scandium (45Sc),
manganese (55Mn), cobalt (59Co), arsenic (75As), yttrium (89Y), niobium (93Nb), rhodium (103Rh), io-
dine (127I), cesium (133Cs), praseodymium (141Pr), terbium (159Tb), holmium (165Ho), thulium (169Tm),
gold (197Au) and bismuth (209Bi).
Note that the atom cannot be considered a geometric sphere with clearly defined radius, diame-
ter, surface area, volume etc. because electron shell parameters are quantum mechanical. Chemistry
thus defines the radius of a given element as half the distance between the nuclei of two neighboring
atoms. Obviously, this depends on the type of chemical bond between these two atoms.
2.3 Conventional parameters of the nucleus 43
272 pm. Atom diameters ∅ thus range from ca. 60 to ca. 600 pm, i.e. approximately 6
to 60⋅10−11 m. Atomic radius is often given in Ångströms (1 Å = 10−10 m = 100 pm), i.e.
approximately 3 to 30⋅10−11 m or 30–300 Å.
nucleus
electron
orbit
1s electrons (2)
2s electrons (2)
1p electrons (2)
Fig. 2.4: Simplified illustration of diameters of the carbon nucleus 12C vs. carbon atom. The
dimension of the nucleus is on an fm scale, which is a factor of about 1,000 smaller than the atom.
The image dramatically overestimates the dimension of the nucleus. For carbon, the radius of the
nucleus is r ≈ 2.1 fm = 2.1⋅10−15 m and the diameter is ∅ ≈ 4.2 fm = 4.2⋅10−15 m. The nucleus of the
carbon atom occupies around 5⋅10−12% of the volume of the whole carbon atom, but collects more
than 99.9% of the atom’s total mass.
λ =
400–750 nm
500 nm
100 nm
< 1 nm
Fig. 2.5: Wavelength of visible light related to the dimension of an atom. Light provides dimensions
much larger than the object of investigation.
However, similar to the dimension of the shell of the atom, it is rather difficult
to exactly quantify the radius of a nucleus. Experimental approaches are needed to
obtain data not accessible via e.g. visible light spectroscopy. Interestingly, sub-
atomic particles are employed instead, such as neutrons (with kinetic energy rang-
ing from En = 14 MeV − 1.4 GeV) and electrons for scattering experiments. These
particles penetrate the atom and are scattered at the nucleus. The nuclei investi-
gated are those of stable isotopes.
Therefore, the question is: what does a nucleus look like?
1. Is it a perfect sphere – or not?
2. Are atomic nuclei spherically shaped – or are they deformed?
3. Are the nucleons distributed homogeneously within the nucleus – or not?
Fig. 2.6: Experimental determination of dimensions rmass and rcharge of an atomic nucleus by
neutron or electron scattering. Experimentally obtained sizes differ; neutron scattering indicates a
larger size than that obtained by electron scattering for one and the same atom.
the radius also depends on the mass number A. This relation assumes the nucleus
is at ground state energy and has a nondeformed, spherical shape, i.e. V = 4/3 πr3,
and for its radius, r ≈ A⅓.
To account for the discrepancy between experimentally measured radii (rmass and
charge
r ) and the theoretical radius (r), a proportionality factor is used: the so-called ra-
dius parameter ro. The resulting quantitative correlations between radius and mass
number are given by eqs. (2.2a) and (2.2b). When discussing the nuclear radius, two
different values are used – a mass radius parameter romass and a charge radius parame-
ter rocharge. The mass radius parameter is 1.3 fm ≤ romass ≤ 1.4 fm, depending on the Z
and N content of a given A. It represents the contribution of a single nucleon to the
overall radius of a nucleus composed of a given number of nucleons. Likewise, the
charge radius parameter reflects the contribution of each proton to the overall size of a
nucleus. It is rocharge = 1.23 fm, which is a bit smaller than the mass radius parameter.
1
r ≈ A =3 (2:1)
1
rmass = f ðAÞ = rmass
0 A =3 (2:2a)
1
rcharge = f ðZ, AÞ = rcharge
0 A =3 (2:2b)
Consequently, radii derived for identical A differ. Table 2.2 gives examples of the mass
radius for three mass numbers: a small nucleus (A = 10), a larger nucleus (A = 100),
and the carbon isotope 12C.
One could say that, because the proton is lighter in mass than the neutron (see
Tab. 2.1), the charge radius parameter should be smaller. In fact it is, but for a different
reason; experimental evidence shows that protons are more tightly packed within the
46 2 The atom’s structure II: Nucleons and nucleus
Tab. 2.2: Radii of atoms of mass numbers 10, 12, and 100 calculated using an averaged mass
radius parameters romass = 1.35 fm and charge radius parameters rocharge = 1.23 fm. Note that A = 12
stands for the carbon isotope 12C, but is synonymous with other mixtures (“isobars”) for A = Z + N,
such as 12B (5 protons, 7 neutrons) or 12N (7 protons, 5 neutrons).
A Mass radius: rmass = f(A) = romass A⅓ Charge radius: rcharge = f(Z,A) = rocharge A⅓
center of the nucleus. Consequently, there is a higher charge density eρ defined for
specific volume elements dν of the volume V of the nucleus. The variation of charge
density in terms of ρ(r) yields average radius r½ values. The corresponding relation-
ship utilizes ρo, the ultimate density of matter. Ideally, it would reflect the density of a
single nucleon itself – but expanded into a larger dimension.
ρ0
ρðrÞ = (2:3)
1 + exp r − r1=2 =a
Figure 2.7 illustrates the concept of a constant charge density changing along the
radius of a nucleus. It remains constant (at ρ/ρo = 1) starting from the center of the
nucleus until an outer part of the spherical nucleus is reached. This outer region has
a lower charge density (ranging from ρ/ρo = 0.9 to 0.1, depending on the nucleus).
ρ
ρ₀
edge zone (a)
1.0
0.5 r½
< r² >
0 1 2 3 4 5 6
r (fm)
Fig. 2.7: Charge density distribution within an atom (FERMI distribution of charge density). The ratio
of ρ/ρo on the y-axis is expressed versus distance from the center of the nucleus. Resulting
parameters include: a variable zone, a, of the edge; an average radius r½ (ratio of ρ/ρo = 0.5); and
a radius r derived as an effective radius; r.m.s. = “root mean square” = √ < r2 > .
2.3 Conventional parameters of the nucleus 47
The drop in charge density indicates that, although there is still a constant density of
nucleons, the percentage of protons is decreasing within an edge zone of dimension
a. The ratio of ρ/ρo equal to 0.5 defines the average radius r½. Another noteworthy
value is the effective radius, called r.m.s. = “root mean square” = √ < r2 > . Figures 2.8
and 2.9 illustrate the charge density in terms of relative and absolute units.
ρ
ρ₀
2.5
2.0
1.5
light nuclei
medium-heavy nuclei
heavy nuclei
1.0
Sr
Mg
0.5 V Co
Ca In Sb
H C O Au Bi
1
Hx 10 He
0 2 4 6 8 10
r (fm)
Fig. 2.8: Charge density of selected nuclei, shown on a relative scale (i.e. as the ratio of ρ/ρo = f(r)).
With permission: R Hofstadter, Nuclear and nucleon scattering of high-energy electrons, Ann. Rev.
Nucl. Sci. 7 (1957) 231.
The effective radius, < r2 >, depends on the charge density distribution along a radial
or volume element according to:
ð
∞
0.10
0.10
²⁰⁸Pb
0.10
¹²⁴Sn
0.10
0.10 ⁵⁸Ni
6 8 10
0.10 ⁴⁸Ca 6 8 10
0.10 ⁴⁰Ca
4 6 8 10
charge density (e fm¯³)
0.08
¹²C 4 6 8 10
0.06
0.04 4 6 8 10
⁴He
0.02 2 4 6 8 10
0
0 2 4 6 8 10
r (fm)
Fig. 2.9: Charge density of selected nuclei, shown on an absolute scale (i.e. ρ = f(r)) with units
(e / fm3). With permission: B Frois, in: P Blasi and RA Ricci (eds.), Proc. Int. Conf. Nucl. Phys.,
Florence, 1983, Tipografia Compositoria Bolgna, vol. 2, p. 221.
Precise determination of atomic radii remains a challenge, and not only because
nuclei have no sharp edges/surfaces. Theoretical models are needed to describe the
status of a nucleus in terms of its mass, which is obviously one key parameter that
correlates with radius. However, measuring the mass of an individual nucleus with
extreme precision remains an analytical problem. Moreover, theoretical models and
recent measurements indicate that the shape of the nucleus – even in its ground
state – is not necessarily a sphere. A nucleus may also be deformed in its ground
state. Figure 2.10 illustrates three nuclei of natural elements (silver, gold, and neon)
and their respective deformations: prolate for Ag, oblate for Au, and hexadecupol
for Ne.
The two kinds of nucleons each have an individual absolute mass (see Tab. 2.1),
which can be used to calculate the mass of the nucleus. The nucleus represents al-
most, but not all, of the mass of the atom. The mass of the shell is the mass of the
electrons Z⋅me. In addition, the binding energy EB(e) (n) of all the electrons must be
considered. This will be different for each electron shell because of the different
main quantum number, as discussed in Chapter 1. Overall, the mass of the shell
components does not contribute significantly to the overall mass of the atom, but is
definitely nonzero.11
The mass of a single electron with zero kinetic energy, called “rest mass”, is only me = 9.109382⋅10−31 kg
(Tab. 2.1). The overall binding energy of all the electrons (as estimated via the THOMAS–FERMI model)
is found using eq. (1.2) En = −Ry Z2 ⋅ 1/n2, or more precisely using Be(Z) = 14.4382 Z2.39 + 1.55468⋅10−6
Z5.35 eV.
50 2 The atom’s structure II: Nucleons and nucleus
There is, however, a more practical approach to calculating atomic mass, using
the “unified atomic mass unit” or “atomic mass unit” u. Instead of using the abso-
lute numbers of mass and binding energy, it considers the mass of a stable isotope
of a prominent atom as a reference value. This mass is very precisely known from
experimental observations. The mass is divided by the number of nucleons, provid-
ing the mass contribution of an average nucleon to the whole atom. This approach
has the advantage that several potential unknowns – such as the mass contribution
of the electrons, their binding energies, and the nucleon binding energies – are al-
ready included in this experiment-based approach (see below).
The isotope selected in determining the atomic mass unit is carbon-12, the most
abundant of the two naturally occurring, stable isotopes of carbon (abundances are
98.93% and 1.07% for 12C and 13C, respectively). The experimentally determined ab-
solute mass of one single carbon-12 atom is 19.92648⋅10−27 kg. It is divided by its
mass number 12, i.e. the 12 nucleons, giving 1.66054⋅10−27 kg – this represents the
“atomic mass unit”, u (see Fig. 2.11).
Mnuclide 12
6 C
u= (2:8)
12
12
Fig. 2.11: Composition of the carbon-12 isotope, which serves as the reference atom in calculating
the atomic mass unit, u. The isotope is “symmetrically” composed of the same number of the three
subatomic particles: 6 protons (orange), 6 neutrons (white), and 6 electrons. The “real” weight of
the atom is well known from experiments.
2.3 Conventional parameters of the nucleus 51
With this parameter in hand, the absolute mass of every other isotope is easily
estimated12 by multiplying the mass number A of the given isotope by the atomic
mass unit. In addition, the atomic mass unit may serve not only as a proportionality
factor to calculate the mass of atoms; it also serves as a measure of subatomic parti-
cle mass. For the electron, proton, and nucleon, mass can be expressed as fractions
of u.13 The proton and neutron mass are thus 1.008665 u and 1.007825 u, i.e. slightly
larger than 1. These values reflect the contributions of electron mass, electron bind-
ing energy, and nucleon binding energy to the overall mass of the atom. For the elec-
tron, the value is me = 5.4858⋅10−4 u (see also Tab. 2.1).
2.3.3 Density
Table 2.4 shows the density of four typical metals, determined for atoms packed in
a metallic state. Their densities range from 2.70 to 11.34 kg/m3. If the radius, volume
(see eq. (2.2a)), and mass of a given nucleus (m ≈ u⋅A) are known, it is possible to
calculate the nuclear density as V/m (see eq. (2.9)).
Tab. 2.4: The density of nuclei vs. the density of atoms. The chemical
element aluminum has a densities of or 2699 kg/m3, respectively (metallic
state of the natural elements at 20 °C temperature). For the nucleus 30Al or
aluminum-30, the density is estimated using the mass number A, mass
(in kg) and size (radius and volume) of the nucleus. The resulting density of
the aluminum nucleus is ≈2.8⋅10+17 kg/m3, compared to 2.70⋅103 kg/m3 for
the aluminum atom.
Parameters
However, since every isotope varies in its nucleon interactions, and the electron binding en-
ergy varies (albeit weakly) with Z, calculated and experimental mass may differ slightly.
The atomic mass unit, u, thus becomes a proportionality factor for very small mass values. This
is similar to the AVOGADRO constant used in chemistry to define an ensemble of a large number of
atoms or molecules: 1 mol = 6.022141⋅1023 atoms or molecules. Although unusual, one may express
any large number in molar fractions. Other large-scale units, such as those measuring distance, are
the light-year and the parsec. A light-year is the distance that light travels in vacuum in one year
and is 1016 m. In astronomy, the parsec is used, which is equal to approximately 3.26 light-years.
52 2 The atom’s structure II: Nucleons and nucleus
1 3 mass 3
4
π rmass · A =3
3 π r0
4
V 3 0
ρ nucleus
= ≈ = (2:9)
m u·A u
For aluminum, for example, the calculated nuclear density is ρnucleus ≈ 2.88⋅10+17 kg/m3
compared to 2.70 kg/m3 for the aluminum atom. That is a difference of 17 orders of mag-
nitude! It again speaks to the fact that an atom’s mass is almost exclusively located in
the nucleus, while the nuclear volume is an almost negligible fraction of the whole
atom. The atom in general thus appears to be an almost empty space, which subatomic
particles and photons can easily penetrate.
In fact, the density of each nucleus is more or less the same, because the nu-
cleus density ρnucleus ultimately depends on the relationship between radius and
mass number. With a mean distance of about 1.8 fm between nucleons, there are
about 0.13 nucleons per fm3 volume in the nucleus. The most dense nuclei have
around 0.17 nucleons per fm3. Furthermore, the density of nucleons packed within
an atomic nucleus is the highest density of matter achievable on earth.14
The dimensions of an atom are extremely small (see Chapter 1); the dimensions
of the atom’s nucleus are even smaller. Since the scale of a 10−43 m3 volume or a
10+17 kg/m3 density is hard to picture, let us therefore try to translate those numbers
into a more familiar scale.
A cube of aluminum containing 1 mole of aluminum atoms has a weight of
26.98 g (obtained from the PTE). In the form of a cube, and with the density of alu-
minum of 2.70 g/cm3, the cube’s volume and edge length are 9.99 cm3 and 2.15 cm,
respectively. Suppose that this same cube consisted of aluminum nuclei instead of
atoms; its weight would be an incredible 2.76⋅1012 kg or 2.76⋅109 t! The weight of
the complete Cheops pyramid is of the same magnitude – it weighs about 6⋅106 t.
Why can we not simply add the absolute mass of each neutron and proton15 to de-
termine the absolute mass of a nucleus? The answer is: the (absolute) mass of an
unbound nucleon differs from the (real) mass of a bound nucleon. Fantastically,
m“bound” nucleon ≠ m“free” nucleon. In the case of nuclei, the whole is not the sum of its
parts.
The ultimate density of nucleons packed tightly inside the nucleus at “infinitely diluted state”
according to the drop model of the atom nucleus (see Chapter 3) is ≈2.96⋅10+17 kg/m3.
Masses of the proton and the neutron are from Tab. 2.1 and valid in the case of mass at rest
energy (mo).
2.4 Mass vs. energy of the nucleus: The mass defect 53
Let us for example consider the α-particle again, i.e. the nucleus of the helium
isotope 4He. It consists of two protons and two neutrons (see Fig. 2.2). The experi-
mental value for the mass of the He atom is 4.00260325415 u.
For the He nucleus, the mass defect value is Δm = 4.00150 u – 4.03188 u = −0.030377 u.
This relative difference is not dramatic, being only <1% in the 4He example. However,
the impact of mass defect on the understanding of matter and energy sources is one of
the most dramatic effects of our material world.16
Using the experimental mass of the nucleus, along with the precise rest mass of
the proton and neutron, values of mass defect are tabulated17 for all the stable
nuclei.
To understand this effect takes one to the philosophic question of “ . . . Dass ich erkenne, was
die Welt // Im Innersten zusammenhält, . . . ” JOHANN WOLFGANG GOETHE, Faust: The First Part of the
Tragedy, 1808: “ . . . So that I may perceive whatever holds // The world together in its inmost
folds.”
If not otherwise indicated, data used throughout the first edition of this textbook were from M
Wang et al.: The AME2012 atomic mass evaluation (II). Tables, graphs and references, Chinese Phys-
ics C 36, (2012) 1603–2014. The database is regularly updated, with new editions published in 2016
and 2020. This textbook refers to the 2020 database by M Wang et al., Chinese Phys. C 45 (2021)
030003 (https://iopscience.iop.org/article/10.1088/1674-1137/abddaf/pdf) For nuclides relevant to the
present textbook, data are listed in Table 14.4 with full precision.
54 2 The atom’s structure II: Nucleons and nucleus
Fig. 2.12: Nucleons bound together within the nucleus form a nucleus of an atom. This nucleus is
lighter in weight (and consequently, lower in energy) than the sum of its identical, unbound
components. When converting a number of individual, “free” nucleons into an atomic nucleus, the
nucleus weighs less than the sum of its nonbound components. There is a “mass defect”!
Tab. 2.5: Mass balance and Δmdefect for the 4He nucleus.
= Z⋅mp + N⋅mn − Δm
Example
difference .
Another important expression is the so-called mass excess Δmexcess (or simply Δ). It is
the difference between the “real”, experimentally determined mass of the nuclide
mnucleus (i.e. its “atomic mass”), and the “relative” mass number A of the same nuclide:
Δ is typically expressed in units of the atomic mass unit, u. The mass excess value,
obtained by eq. (2.11), yields energy when multiplied by c2. For 12C (carbon-12), by
definition the value of Δ in eq. (2.11) must equal 0, according to eq. (2.8). While the
2.4 Mass vs. energy of the nucleus: The mass defect 55
mass defect has an obvious physical meaning (as the equivalent binding energy
within an atomic nucleus), the mass excess is to be regarded as a useful auxiliary
computational quantity.
Some nuclides have a “positive” mass excess relative to carbon-12 (all stable
isotopes with A < 16), while others have a “negative” mass excess (see Tab. 2.6). The
value of Δ ranges from about +16 MeV to − 90 MeV. Figure 2.13 illustrates the trend
in mass excess relative to proton number, Z.
+ 20
0 1
- 20
Δexcess (MeV)
- 40
- 60
- 80
- 100
0 10 20 30 40 50 60 70 80 90
Z
Fig. 2.13: Correlation between mass excess and the proton number Z of stable nuclei. [1] carbon-12
reference line.
Nevertheless, the key to understanding nucleon binding lies in the precise correla-
tion between absolute atomic mass and mean nucleon binding energy. Both mass
defect Δm and mass excess Δ are intermediates in this direction.
Where does that mass, “the mass defect Δm”, go? Of course, mass cannot disappear –
it is converted to energy. Obviously, there is a reason for nucleons to bind together:
in doing so, they reach a lower final energy state compared to the initial (unbound)
state. This amount of energy, ΔE, is equivalent to the mass defect Δmdefect, which ap-
pears to be “lost”:
56 2 The atom’s structure II: Nucleons and nucleus
ΔE = Δmdefect c2 (2:12)
Once the nucleons are within a very small distance of each other (on the order of
fm, i.e. the dimension of the atom nucleus), they are attracted to each other by a
“strong force” – the strongest force known in our universe. The energy saved (– ΔE)
when nucleons bind together, compared to their unbound state, is called “overall
binding energy”. The equivalents of Δmdefect and ΔE thus reflect the overall binding
energy, EB, of the nucleus. Accordingly, EB ≈ A. EB corresponds not only to mass in
general, but also to the number of protons and neutrons, and is given by EB = f(Z,
N)A=constant (see Chapter 3).
Returning to the example in Tab. 2.5, the overall binding energy of the 4He nu-
cleus is 0.03038 u = 0.05045⋅10−27 kg in terms of mass and 4.53⋅10−12 J (or 28.295
660 MeV) in terms of energy.18 The resulting average or mean binding energy per
nucleon is ĒB = EB /A. The binding energy per nucleon within the 4He nucleus (where
A = 4) is thus ĒB = 7.073 916 MeV.
^
EB = ΔE (2:14)
EB
EB = (2:15)
A
How are binding energies calculated? Nucleon binding energy correlates with mass
defect via eq. (2.12) and (2.14), (i.e., EB = ΔE = Δmdefectc2), and with mean binding en-
ergy via eq. (2.15). Let us calculate the mean binding energy of the 12 nucleons
within a carbon-12 nucleus. The absolute experimental mass of the atom 12C is
12.0000 u = 12 × 1.66054⋅10−27 kg = 19.92648⋅10−27 kg. Subtracting the atomic mass
and binding energy of 6 electrons, the mass of the 12C nucleus becomes 11.99676
u. On the other hand, the sum of 6 protons and 6 neutrons (in u) is 12.0957 u. Thus,
the mass defect becomes Δmdefect = 11.996761 u – 12.0957 u = −0.098940 u. In terms
of energy (1 u = 931.494 MeV), it becomes ΔE = 92161.728 MeV. This is the nucleon
binding energy EB of the whole nucleus. Finally, the mean nucleon binding energy
ĒB is EB divided by 12:
12 EB 6mp + 6mn − m 12 6 C 931.494 MeV
EB 6 C = = = 7.680 MeV (2:16)
12 12
Compared to the mean nucleon binding energy of the 4He nucleus, ĒB(4He) =
7.074 MeV, the 12 nucleons of the 12C nucleus are bound more strongly. Table 2.6
lists for a light nucleus (4He), for 12C, for a medium mass number nucleus (56Fe),
and a very heavy nucleus (238U) the values of experimental atomic mass, mean
binding energy, and mass excess.
Tab. 2.6: Experimental atomic mass, mass excess and mean binding energy of
4
He, 12C, 56Fe, and 238U.
Overall binding energy ranges from about 28 MeV for 4He to about 1800 MeV for
238
U. Maximum values for mean nucleon binding energy are ĒB = 8.790 MeV for
56
Fe, 8.792 MeV for 58Fe, and 8.794 MeV for 62Ni. Interestingly, the ĒB values are
very similar compared to the strongly varying mass numbers and atomic weights, at
least for most nuclei of A > 10. In this range of mass numbers, average values for ĒB
are 8.18 ± 0.62 MeV for the ca. 300 stable and more than 3000 unstable nuclei. The
appendix contains a table (14.4) listing relevant “atomic mass data” for the nuclides
discussed throughout this textbook. It should be noted that nuclear data compila-
tions prefer to report mass excess, Δ, instead of mass defect, Δm. These terms corre-
late according to eq. (2.17):
Δm = N · mn + Z · mp − A · u − Δ=c2. (2:17)
2.5 Outlook
With the numbers discussed above, one may summarize the rather classical param-
eters of an atom’s nucleus, i.e. charge, size (radius and volume), mass, and density.
An individual atom’s size and weight are extremely small, and thus they appear to
be invisible and undetectable by the human senses. Nuclei are even smaller than
atoms, by several orders of magnitude. Nuclear mass and volume can be converted
into an impressive, unimaginable density. The contributions made by philosophers
more than 2000 years ago and by scientists about 150 years ago regarding existent
but invisible atomic and subatomic particles earn great respect!
58 2 The atom’s structure II: Nucleons and nucleus
However, in the case of atoms, size is not nearly as important as mass and en-
ergy, particularly of the atomic nucleus. How mass (via nucleons) is arranged
within an atomic nucleus not only affects the characteristic energy of that nucleus,
but also its stability or instability. Chapter 3 continues with the binding energy of
an atomic nucleus and how this experimental parameter challenges theoretical
models of nuclear composition.
Finally, in addition to the “conventional” properties of atomic nuclei such as
mass and radius, there are far more individual parameters. These were identified
within the last century when quantum physics entered the field. The relevant identi-
ties are spin, isospin, electric moment, magnetic moment, and parity (see Tab. 2.7),
which will be introduced in the following chapters. They are relevant to the funda-
mental processes of transforming an unstable nucleus into a stable one, and for the
process of energetic de-excitation of nuclear states into less energetic (i.e. more stable)
states.
Aim: Although it is energetically expensive for two nucleons to approach each other,
they are ultimately able to create a bound state with a “win” in energy relative to the
unbound nucleons. Above all, it is the enormous energy density that makes the nu-
cleus of an atom an extraordinary object of research. The total and mean nucleon
binding energies (EB and ĒB) are key parameters in nuclear science. The amount of
energy a nucleus wins depends on the number of nucleons involved, A.
To understand the systematics of stable (and unstable) nuclides, nuclear sci-
ence structures mass numbers according to the number of protons and neutrons.
The Chart of Nuclides correlates neutron number N with proton number Z in a Car-
tesian coordinate system, revealing how parameters of stability (and processes of
nuclei transformation) relate to both mass number and varying Z + N combinations.
The localization of the stable nuclides and their experimental mass values along
a set of parameters (A, Z, N) results in a correlation of type ĒB = f(A, Z, N). The classi-
cal theory used to explain this correlation utilizes the “liquid drop model” (LDM) of
the atomic nucleus, postulating that all protons are identical and all neutrons are
identical, and all nucleons are distributed homogeneously within the nucleus – like
H2O molecules within a drop of liquid water. The semi-empiric mathematics quantify-
ing this experimental dependency is the so-called WEIZSÄCKER equation.
Within the existing sets of (A, Z, N), there is a surprising overexpression of a few
numbers: namely 2, 8, 20, 28, 50, 82 (both for Z and N), and 126 for N only. Nuclei
expressing these numbers of protons and/or neutrons are generally more stable than
predicted by the LDM. Consequently, another theory accompanies the liquid drop
model theory: the “nuclear shell model” (NSM). Similar to the orbital theory of elec-
trons, both protons and neutrons are assumed to exist at characteristic shell levels
with specific quantum numbers.
Atomic hydrogen is defined by the one proton in its nucleus and the one electron in
its shell. Isotopes of hydrogen originate from an increasing number of neutrons (see
Fig. 2.2). Nucleon mixtures for hydrogen are thus composed of proton-to-neutron ra-
tios of 1:0, 1:1 (deuterium), 1:2 (tritium), etc. However, only two of the mixtures result
in a stable isotope, namely at proton-to-neutron ratios of 1:0 and 1:1, i.e. the 1H and
2
H isotopes. The same is true of helium; the helium isotopes exist as combinations of
its two protons with 1 neutron (3He), 2 neutrons (4He), or more neutrons (5He, 6He,
7
He, etc.). Only two of these mixtures guarantee a stable nucleus, namely 3He and
4
He. Likewise, lithium has only two stable isotopes: 6Li and 7Li. For the next chemical
element, there is only one stable isotope: 9Be.
https://doi.org/10.1515/9783110742725-003
60 3 Nucleons: Binding energy and shell structure
A Cartesian coordinate system, where the x-axis represents the number of neu-
trons N and the y-axis shows the number of protons Z, creates the Chart of Nuclides.1
It starts with the lightest nuclide at Z = 1 and N = 0, i.e. with the hydrogen isotope 1H.
(From an abstract point of view, one could also start from Z = 0 and N = 1, represent-
ing an unbound neutron instead of a nuclide.) Figure 3.1 shows this correlation for
the stable isotopes of the first four chemical elements. It demonstrates that for these
elements, only a few possible combinations of protons and neutrons result in stable
nuclides. Isotopes like 3H, 4H, 2He, 5He, 6He, 3Li, 4Li, 5Li, 8Li, 9Li, etc. do exist – but
are not stable! These will appear later in this textbook, starting in Chapter 6.
9Be
4
100
6Li 7Li
3
7.59 92.41
Z
3He 4He
2
0.000134 99.999866
1H 2H
1
99.9985 0.0115
1n
0
1 2 3 4 5 6
N
Fig. 3.1: The Chart of Nuclides plots the proton number (Z) versus the neutron number (N) for stable
nuclides on a Cartesian coordinate system (x, y). Stable nuclides are typically labeled in black,
alongside their mass number A. This figure is a simplified excerpt from the Chart of Nuclides,
showing the stable isotopes of the four lightest elements: hydrogen, helium, lithium, and
beryllium. The lower number in each field represents the natural abundance (%) of that isotope
within a given element. This means hydrogen has a natural abundance (naturalH) of 99.9985% (1H)
and 0.0115% (2H), totaling 100%. Note that the (1, 0) combination representing a single, unbound
neutron is in gray: it is not stable.
The PTE and the Chart of Nuclides thus represent two different perspectives on chem-
ical and nuclear sciences (see Fig. 3.2). The PTE predicts the chemical behavior of an
element and reveals the system of electron distribution within the inner structure of
the atomic shell. It is a periodic system, despite the fact that chemical elements in the
PTE (generally) appear in order of increasing mass. The main information is on the
chemistry of an element E following the number of electrons, which in turn are repre-
sentative of Z.
In the following, the word “nuclide(s)” is preferred when referring to the whole unstable atom,
rather than the nucleus alone. Isotopes are nuclides of one and the same chemical element.
3.1 Stable nuclides and the Chart of Nuclides 61
Fig. 3.2: The main parameters used by the Periodic Table of Elements and the Chart of Nuclides.
The Chart of Nuclides is a coordinate system with the number of neutrons and protons
on the x- and y-axes, respectively. The y-axis ranges from Z = 1 (representing the hydro-
gen isotopes) to Z = 118, encompassing all known chemical elements, stable or unstable
(at least so far). The x-axis scales from N = 0 (the hydrogen isotope 1H) to N > 170 for the
artificially produced super-heavy, transactinide elements. This results in a matrix of
about N × Z = 170 ×120 = 20,400 possible combinations. Yet, little more than 1% (281)3 of
Chemists typically handle stable isotopes and may not need the Chart of Nuclides. Likewise, nu-
clear scientists interested in the properties of an atom’s nucleus or in nuclear transformations may
not look at the chemistry of the nuclides involved. Nevertheless, many scientific research topics
require both the chemical and nuclear parameters of unstable nuclides.
The total number of stable nuclides varies between textbooks. This is due in part to the some-
what philosophical question: what is stability? If “absolute” stability is meant, then with increasing
Z, the “last” stable nuclides are those of lead (Z = 82) and bismuth (Z = 83): 206Pb, 207Pb, 208Pb and
62 3 Nucleons: Binding energy and shell structure
all the theoretical combinations are “realized” on earth as stable nuclides. Within the
Chart of Nuclides, the black squares denote stable nuclides.
Abstractly, the squares can represent the surface of a box which on all six faces
may offer characteristic information. At the surface, which is depicted in the Chart
of Nuclides, there are data on the isotope notation, the abundance of a stable iso-
tope within a given element, and the cross-section σ of the most relevant nuclear
reaction4 process.
Figure 3.3 illustrates the Chart of Nuclides summarizing the stable isotopes
(black squares). Four reference lines are marked. Line one represents nuclides with
Z and N symmetry (N = Z).
Three more lines indicate situations where one parameter is constant (“iso”-)
among the values of Z, N and A. Line 2 characterizes the isotopes, i.e. nuclides where
Z = constant and (N, A) are variable. Line 3 shows the same for neutrons, called an
isotone, i.e. N = constant and (Z, A) are variables. Line 4 marks nuclides of constant
mass number A, called an isobar, i.e. A = constant and (Z, N) values vary. Lines of
type 1 (i.e. when N–Z = 0 = constant) may turn into N–Z > 0, but again N–Z must be
aconstant value. Lines of this type are referred to as isodiapheres (see Tab. 3.1).
The value of A denotes the “theoretical” mass. However, “real” mass differs from this
value because of the mass defect. We are ultimately interested in the real (e.g. experi-
mental) mass of individual nuclides because mass is tantamount to energy. Only
knowledge of the precise mass of a nucleus can open access to the overall nucleon
binding energy. Finally, knowledge of the overall and mean nucleon binding ener-
gies is the key to correlating experimental values with theoretical models, ultimately
to address the question: how and why do nucleons exist inside an atomic nucleus?
209
Bi. Beyond bismuth, chemical elements only have (more or less) unstable isotopes. In terms of
“relative” stability, nuclides may be considered whose half-life (t½) is greater than the age of our
planet: about 4.5⋅109 years (see Chapter 5). Translated into a human scale, these isotopes decay by
a negligible fraction (<0.00000178%) over the average life-span of 80 years. Over the observation
period of a human, such a nuclide is “quasi-stable”. This is true even referenced against the age of
mankind, about 2⋅106 years. Over this period, an isotope of t½ > 4.5⋅109 years will have changed
by 0.044% – representing de facto stability. However, even those nuclides having half-lives of sev-
eral thousand years – the thorium isotope 232Th (t½ = 1.405⋅1010 years) and the three uranium iso-
topes (234U, 235U and 238U with t½ = 2.455⋅105, 7.038⋅108 and 4.468⋅109 years – are not absolutely
stable. Nevertheless, in the Chart of Nuclides they are denoted by (partially) black fields.
See Chapter 13 for nuclear reactions.
3.2 Mass, mass defect, nucleon binding energy 63
80
4 3 1
60
Z 40 2
20
0
0 20 40 60 80 100 120
N = A-Z
Fig. 3.3: The Chart of Nuclides indicating the stable isotopes (black squares). Overall, the nuclides
do not follow a straight line of type Z = N (line 1). Line 1 is followed only for mass numbers <40, i.e.
proton and neutron numbers <20. Above Z = 20, the distribution is in favor of neutron excess
(N > Z). Three more lines are indicated, each marking the line of a certain constant (“iso”-)
parameter among the variations of Z, N, and A. Line 2 characterizes the isotopes, i.e. nuclides of
Z = constant, with variable N and A values. Line 3 shows an isotone, where N = constant and (Z, A)
are variables. Line 4 marks nuclides of constant mass number A, called an isobar, i.e. A = constant
and Z and N varying. Isodiapher lines (of N–Z > 0 = constant) are not shown.
Isodiaphers N − Z = constant, (N = Z) e.g. H, He, Li, B, C, N, O, . . .
But first, what is meant by “precise mass of a nucleus”? It requires that the real mass
of an atom’s nucleus should be experimentally determined as precisely as possible.5 This
There are several methods to do so, with laser spectroscopy being the most recent and accu-
rate. The accuracy of mass determination today is comparable to weighing the mass of a passen-
ger airplane, such as an Airbus 380–800, and detecting whether 1 out of >600 passengers carries
64 3 Nucleons: Binding energy and shell structure
is practically achievable in the case of stable nuclides.6 So, what information is included
in the experimental database of stable nuclides? This database includes sets of nucleon
numbers in terms of three parameters (Z, N, and A), as well as experimentally determined
atomic masses. The number of protons runs from Z = 1 (representing the chemical ele-
ment hydrogen) to Z = 83 (bismuth) and includes a few very long-lived isotopes at Z = 90
and 92 (thorium and uranium). The number of neutrons covers a larger range, starting
from N = 0 (corresponding to 1H) and approaches N = 146 (for the heaviest “relatively”
stable isotope of uranium, 238U). The mass number A covers a range of 1 ≤A ≤238.
This is the database to develop a theory about the nucleon interactions respon-
sible for mass defect and overall nucleon binding energy. As previously discussed,
the overall binding energy can be normalized to the number of nucleons in each
nucleus, resulting in the mean nucleon binding energy.7 It is interesting to see how
the mean nucleon binding energy varies with increasing mass number A. Figure 3.4
arranges this correlation according to ĒB = f(A). Several observations and conclu-
sions can be drawn from this correlation; these are listed in Tab. 3.2.
With the experimental database in hand, basic theories about the atomic nu-
cleus relying on the number of nucleons and the experimental mass of a nucleus
can be addressed (see Tab. 3.3).
There are two principal models to explain the observations and the additional
facts in Tab. 3.2 and the questions in Tab. 3.3: the “liquid drop model” (LDM) and
the “nuclear shell model” (NSM).8 The main features of these two models are sum-
marized in Tab. 3.4.
The LDM postulates that all protons are identical and all neutrons are identical:
one proton cannot be distinguished from another, the same holding true for neu-
trons. It further posits that the nucleons are distributed homogeneously throughout
the nucleus (similar to the early RUTHERFORD plum pudding model of the entire
atom, which expected to find protons and electrons distributed throughout the
atom). All these identical nucleons distributed homogeneously within the nucleus
should behave like H2O molecules within a drop of liquid water. The nucleons rep-
resent an ensemble of identical particles equidistance from each other. This dis-
tance is fixed and cannot be modified by e.g. trying to deform the “liquid drop”. In
fact, it is impossible to compress the liquid drop-like atomic nucleus. It may only
change its shape, while the volume remains constant – like a drop of liquid water.
a single one-cent coin or not. Suppose the airplane weighs 560 tons and the coin 1 g; the precision
is of Δm/m ≤10−9.
Stable isotopes are much easier to analyze than unstable, radioactive ones.
Table 2.6 lists these numbers for four selected nuclei. Standard publications in the nuclear scien-
ces list those and other parameters for almost all the nuclides; see Table 2.7.
Both concepts appeared in the 1930s. These concepts have been verified, yet still today there is
an ongoing process to further qualify these theories.
3.2 Mass, mass defect, nucleon binding energy 65
10
2
9
3 4
8
5
7
6
ĒB [MeV]
5
1
4
0
0 20 40 60 80 100 120 140 160 180 200 220 240
mass number A
Fig. 3.4: Mean nucleon binding energy ĒB as a function of mass number A, shown for the 281 stable
nuclei. ĒB values range from about 1–9 MeV per nucleon. (1) Light nuclides show low values of ĒB.
(2) At A > 50 until about 90, the ĒB values become relatively similar (about 8.18 ±0.62 MeV). (3) The
maximum of ĒB appears around A = 50 (where ĒB = 8.7 to 8.8 MeV), and subsequently the values
decrease. (4) Very heavy nuclides of thorium and uranium (A = 234, 235 and 238, 7.570 MeV for
238
U). (5) Value of ĒB for 12C (ĒB = 7.680 MeV).
Tab. 3.2: Experimental observations from Fig. 3.4, useful for understanding the distribution of
stable nuclei in the Chart of Nuclei and for theoretically explaining correlations of type ĒB = f(A) of
stable nuclei. Some additional statistical facts needed to understand the WEIZSÄCKER equation are
also provided.
. There is a general trend in mean nucleon binding energy versus mass number A. ĒB first
increases from light nuclei to medium mass nuclei, stays relatively constant between ca. < A
< , reaching maximum values for isotopes of iron, cobalt, and nickel and finally decreases
slightly with further increasing mass numbers.
. Values of ĒB are relatively similar – at least when mass number A > . Except the two stable
nuclei H and He (ĒB = . MeV and . MeV, respectively), mean nucleon binding energy
ranges from about to . MeV. This is a relatively narrow range for atomic nuclei of mass
numbers spanning about two orders of magnitude (He to U).
66 3 Nucleons: Binding energy and shell structure
. Relative to the “reference” nucleus C (ĒB = . MeV), there are nuclei with lower mean
nucleon binding energy in the mass range of –, but also nuclei with higher binding energy
per nucleon for mass number A ≥, while most of the nuclei have higher mean nucleon
binding energy.
. The stable nuclei do not distribute within the Chart of Nuclei along a straight line of type Z = N.
This is true only for mass numbers <. At A > , the distribution is in favor of neutron excess
(N > Z).
. In the low mass region, there are several “peaks” along the virtual “line”. Obviously, some
mass numbers create higher values for ĒB than “usual” or “expected”.
. When proton and neutron numbers of a nucleus are counted separately, either even or odd
numbers for Z and N appear. Mass numbers A thus reflect combinations of type (Z, N) of (even,
even), (even, odd) or (odd, even) and (odd, odd). For the stable nuclei, those combinations are
not equally abundant. The (even, even) nuclei dominate and the ratios among the four possible
combinations are about :::. Thus, there are about times more stable nuclei
representing (even, even) numbers compared to (odd, odd) versions.
. The ranges of proton and neutron numbers for stable nuclei are < Z < and < N < ,
respectively. One may expect that the complete range is found more or less balanced within
the combinations reflecting the range of mass numbers, i.e. < A < . However, statistics for
the abundance of numbers of Z and N for stable nuclei reveal an unusually high abundance of
the following six numbers. These numbers are , , , , and for Z; and , , , ,
, and for N. These numbers are called “magic”.
. In contrast to the high abundance of these numbers (, , , , , and ), numbers
corresponding to a magic number ± (e.g. N − or N +) appear surprisingly seldom.
Tab. 3.3: Questions concerning the atomic nucleus and the mean binding energy of its nucleons.
What model best explains the difference between theoretical and experimental mass – the mass
defect?
What theory explains how much energy is needed to make a nucleus out of individual nucleons –
and how much energy is released when atomic structures are transformed into free nucleons?
What theory can describe and ultimately predict the combinations of type (N, Z) reflected in the
structure of a nucleus?
What theory can explain and predict the transformation processes from an unstable nucleus to a
stable one?
3.2 Mass, mass defect, nucleon binding energy 67
Since the protons are distributed homogeneously, charge distribution within the
nucleus is assumed to be uniform.
Within the LDM, one would expect that the specific Z and N numbers do not
matter – it is all about the overall nucleon number, A. Cataloguing the stable nu-
clei, however, reveals a surprising overexpression (in terms of their statistical abun-
dance) of certain, limited numbers for (interestingly) both Z and N. Namely, these
are 2, 8, 20, 28, 50, 82 and (for N only) 126. Because these protons and/or neutrons
numbers are more commonly expressed in stable nuclei than other values of Z and
N, they seem to be a source of higher stability of a nucleus. In the absence of a the-
ory to explain this overexpression, the numbers were called “magic”.
This oddity is addressed by the nuclear shell model theory. The nuclear shell
model does not contradict the LDM, but rather accompanies it. It correlates experi-
mental evidence on the increased number of stable nuclei having “magic” nucleon
numbers with the internal structure of nucleons within the nucleus.
Similar to the shell model used for electrons, protons and neutrons are under-
stood to exist at characteristic shell levels with individual quantum number and a
resulting numbers of nucleons per shell. It applies the concept of quantum numbers
introduced for electrons to nucleons. Similar to how each electron of an atom is dif-
ferent (i.e. each is unique in its set of quantum numbers), now each nucleon is
treated “individually” as well. As an example, in the stable isotope gold-198, all 79
protons and 119 neutrons are different – each of the 198 nucleons is unique.
68 3 Nucleons: Binding energy and shell structure
+
UME
PAIRIING
VOLU
Z2 ((N-Z))2 LDM
EB = LDM A – LDM A⅔ – LDM – LDM ± ¾ (3.1)
A⅓ A A
SYMMETRY
OULOMB
URFACE
PAIRING
AS
SU
CO
The basic equation considers five different terms, although LDM theory is still being
developed and rethought today. Equation (3.1) reproduces a rather standard ver-
sion, while an expanded version is given, for example, in the Handbook of Nuclear
Chemistry.10
The equation may be divided into several parts, also listed in Tab. 3.5. Each
term of this equation has a physical rationale, describing the various ways the two
different types of nucleons contribute to the binding energy. Some terms depend
exclusively on mass number A, while others reflect the individual contributions of
either protons or neutrons. Finally, each term has a coefficient, which overall re-
sembles a polynomial equation. The absolute values of the five coefficients – αLDM,
βLDM, γLDM, ζLDM, and δLDM – are fitted to the trend in “experimental” values of
mean nucleon binding energy in Fig. 3.4.
. . . although a number of other great scientists have contributed to its conception and modifications,
such as BETHE, GAMOV, BOHR AND FERMI. Equation (3.1) is also called the BETHE–WEIZSÄCKER equation.
A Vértes, S Nagy, Z Klencsár, RG Lovas, F Rösch (eds.), Handbook of Nuclear Chemistry, second
edition, Springer 2011.
3.3 The liquid drop model 69
Symbol Sign
Volume A αLDM
+
Surface A⅔ βLDM −
Coulomb Z/A⅓ γLDM −
Asymmetry (N − Z)/A ζLDM −
Pairing /A¾ δLDM +, − or
3.3.1 Volume
Among all the terms listed for eq. (3.1), the “volume term” is the principal one, and
the one showing per se a positive sign. It is positive because energy is gained when
the nucleons are “bound”;11 bounded nucleons have lower potential energy than
unbounded ones. Figure 3.5 depicts an illustration of protons and neutrons within a
nucleus. The more interactions per individual nucleon, the larger the gain in energy
for that particular nucleon.
This increase in mean nucleon binding energy depends exclusively on A ac-
cording to a direct proportionality. The term does not depend on Z nor on N.
Consequently, larger nuclei have higher overall binding energies. The volume
of a sphere is 4/3πr3 and volume ≈ number of nucleons A.12
The volume term expresses EB as proportional to A, regardless of the composition
of A in terms of Z and N. Because A is a unitless number (for stable nuclei ranging
from mass number 2 to about 200),13 the coefficient αLDM is also needed to introduce a
unit: MeV. According to recent modifications of the WEIZSÄCKER equation, its value is
αLDM = 15.15 MeV. When only the volume term of the LDM is considered, the expression
would be EB = αA. The hydrogen isotope 2H and uranium isotope 238U thus should
have overall nucleon binding energies of EB = 30.30 MeV and EB = 3605.70 MeV,
respectively.When expressing these values in terms of mean nucleon binding en-
ergy, each value of EB is divided by A; consequently, for both nuclei, the result is
It reflects the strong interaction of nucleons, mediated by the strong force (see Chapter 7). This
strongest force of our universe attracts nucleons approaching each other. The range of this strong
nuclear force is extremely small, similar to the dimension of nucleons (i.e. at fm scale).
Relationships between size and radius of an atomic nucleus have been discussed in Chapter 2.
Note that the range is not A ≥ 1, but A > 1. The hydrogen isotope 1H cannot be considered as
there is only one nucleon; there are no additional nucleons with which the proton can interact.
70 3 Nucleons: Binding energy and shell structure
1:6
Fig. 3.5: Volume term: protons (orange) and neutrons
(white) within a nucleus attracted by the strong force.
In this simplified two-dimensional drawing of a circle
instead of a sphere, each nucleon within the volume of
a sphere is surrounded by six other nucleons. In
reality, i.e. in three dimensions, there are more.
ĒB = 15.15 MeV. This would be true for all stable nuclei. Figure 3.10 plots the result
of the LDM-based mean nucleon binding energy as ĒB = f(A). The horizontal line
at 15.15 MeV is the resulting hypothetical expression for the volume term only.
3.3.2 Surface
1:4
The surface term calculates the fraction of these nucleons, and varies by the ratio
between volume and surface of a sphere. The surface area of a sphere is S = 4πr2 and
is proportional to A⅔. The surface term thus reads as −βLDMA⅔. Note that the sign is
negative, so this term subtracts a certain value from the volume term.
The mean nucleon binding energy corrected this way is given in Fig. 3.10. Clearly,
this type of correction is most dramatic for light nuclei.14 Consequently, the surface
term correction dominates for nuclei of about A < 10. For increasing A, the ratio be-
tween volume and surface (i.e. of nucleons inside the volume and nucleons at the sur-
face) becomes smaller. Thus, the fraction of nucleons that are more weakly bound (at
the surface) increases with A. The value of (volume term − surface term) for A = 50 be-
comes 11.823 MeV, and approaching heavy nuclei it further decreases. For 238U, the
mathematical outcome for ĒB is 12.578 MeV, instead of 15.15 MeV for the volume term
alone.
Both the volume and surface terms of the liquid drop model take into account the
total number of nucleons, A. However, the additional impact of the positively
charged nucleons must be considered. This is captured in the next two terms.
The first is the “coulomb term”. It realizes that protons (unlike neutrons) un-
dergo coulombic interactions. Protons “classically” should try to achieve a distance
between each other to minimize electrostatic repulsion. Protons are thus distributed
within the nucleus in a way that minimizes repulsion (see Fig. 3.7).
Overall, coulombic interactions tend to lower the overall nucleon binding energy.
While the volume is represented mathematically as proportional to A⅓, coulombic
interaction depends on Z2. In more detail, the coulombic energy is ECOUL = 3/5Z2e2/r,
where e represents the elementary charge unit and r is the radius of the nucleus
(r = ro A⅓). ECOUL thus is proportional to Z2/r and with r proportional to A⅓ it be-
comes a proportionality to Z2/A⅓. The resulting coulomb term is –γLDM Z2/A⅓. The
message is that the more protons there are within a given volume, the more dra-
matically nucleon binding energy is affected.15 However, the absolute amount
The larger the volume of a sphere, the smaller the relative fraction of nucleons at the surface
becomes relative to the overall number of nucleons. For example, a light nucleus such as 2H con-
tains all its nucleons at the “surface”.
Example: The reference nucleus 12C is composed of 6 protons and 6 neutrons. The constituents
of the coulomb term are Z2 = 6 ×6 = 36 and A⅓ = 12⅓ = 2.28943. In this case, Z2/A⅓ equals 15.724. Its
isobar with 7 protons (where A = 12 and N = 5) is 12N. Here, since Z2 = 7 ×7 = 49, the term Z2/A⅓ be-
comes 21.403. The difference in Z2/A⅓ is 36%. This term decreases the 12N nucleon binding energy
(until now exclusively determined by volume and surface term). As a result, 12C is stable, while 12N
is not. In general for a given mass number, the larger the proton excess, the lower the remaining
nucleon binding energy.
72 3 Nucleons: Binding energy and shell structure
Fig. 3.7: Coulomb term: protons (orange) are distributed in a way that minimizes coulombic
repulsion. In this simplified sketch, there are no protons in close contact with each other (left).
Because there is no “empty space” within a typical nucleus,16 direct contact between protons is
avoided by “buffering” with neutrons (right).
subtracted from the volume and surface term is relatively low, as the coefficient
γLDM is only 0.665 MeV.
Figure 3.10 illustrates the impact this third term has on the trend in mean nu-
cleon binding energy. Its effect is more pronounced for large mass numbers. For
A = 50 and A = 238, for example, values calculated so far via volume and surface
term of 10.823 and 12.578 MeV are reduced to 9.076 and 8.762 MeV, respectively.
The database for light stable nuclei indicates that whenever one goes from one
chemical element to the next heavier one (i.e. from Z to Z +1), the number of neu-
trons follows in the same manner. There are several prominent examples among
the light nuclei, such as 2H (1 proton + 1 neutron), 4He (2 protons + 2 neutrons), 6Li
(3 protons + 3 neutrons), . . ., 12C (6 protons + 6 neutrons), . . ., 40Ca (20 protons + 20
neutrons), etc. These nuclei are “symmetric” in the sense that Z = N. However, ex-
perimental database reveals that this is true for a very limited number of stable nu-
clei only. When A > 40, this symmetry is lost. All stable nuclei at Z > 20 and N > 20
tend to deviate from this symmetry (line 1 in Fig. 3.3). The relationship between Z
and N becomes asymmetric (Z ≠ N, or more precisely: N > Z). Asymmetry also means
that whenever protons and neutrons are taken artificially as pairs, one or more neu-
trons remain solo (see Fig. 3.8). This “excess” of neutrons over protons increases
with increasing mass number A. For heavy stable nuclei (such as 238U), the excess
There are exceptions to the rule for several light nuclei, called halo nuclei.
3.3 The liquid drop model 73
For 238U, pairing the 92 protons with neutrons would altogether give 184 nucleons. However,
the mass number is 238, i.e. the remaining 54 nucleons (238–184 = 54) are “excess” neutrons.
For symmetric nuclei, N − Z becomes zero and the entire term vanishes. Increasing the excess of
neutrons over protons creates numbers such that N − Z is >0. For example, the stable fluorine nu-
cleus 19F has 10 neutrons minus 9 protons, making N − Z = 1 and (N − Z)2 = 1. Likewise, 238U has 146
neutrons minus 92 protons, meaning and N − Z = 54 and (N − Z)2 = 2916. The neutron excess alone
results in the 238U nucleus being more than three thousand-fold stronger than 19F. However, in the
end, that square of neutron excess is less impactful, since the asymmetry term relates the neutron
excess to the overall number of nucleons. For 19F, the expression (N − Z)2/A is 1/19 = 0.05263168 MeV;
for 238U it is 13.65126 MeV. This is still a significant difference of more than two orders of magnitude.
74 3 Nucleons: Binding energy and shell structure
3.3.5 Pairing
The final term of the original WEIZSÄCKER equation is the pairing term. In this case,
protons and neutrons are counted separately, resulting in an odd or even proton
number, and an odd or even neutron number. In terms of A = Z + N, the possible com-
binations of (Z, N) nuclei are (even, even), (odd, odd), (even, odd), and (odd, even).
Any mass number A thus reflects one of these four combinations. The question, nev-
ertheless, is: why should one deal with this kind of statistic? The answer refers to the
experimental database (see point 6 in Tab. 3.2). Strikingly, there is a very strange sta-
tistic regarding the stable nuclei. The (even, even) nuclei dominate and the ratios
among the four possible combinations (even, even), (even, odd), (odd, even), and
(odd, odd) are about 30:10:10:1. There are about 30 times more stable (even, even)
nuclei compared to (odd, odd) versions.19 Table 3.6 lists the statistics.
Tab. 3.6: Statistics of (Z, N) combinations of all the stable nuclei in terms of
even or odd proton or neutron numbers and sign of δLDM, the coefficient of
the pairing term in the WEIZSÄCKER equation.
Mathematically, the pairing term depends on the mass number according to 1/A¾.
Pairing further “corrects” the calculated mean nucleon binding energy and is effi-
cient over the whole range of mass number A (see Fig. 3.9).
Two protons (or similarly, two neutrons) may be considered to form a pair in-
side the nucleus.21 From the dominating fraction of (even, even) nuclei in the uni-
verse, it is concluded that (even, even) combinations result in increased stability,
i.e. higher nucleon binding energy. The coefficient δLDM of the pairing term thus
gets a positive sign.
This is mirrored even in the presence of the chemical elements in the universe.
20 63 Li3 , 10 14 50 176 180
5 B5 , 7 N7 , 23 V27 , 71 Lu105 , 73 Ta107 .
For the WEIZSÄCKER equation, the pairing term is rather semi-empirical in character. At this
stage, its physical meaning remains uncertain; here, it simply reflects the status of the database. A
rational explanation will arise only later, in the context of quantum mechanics; see the nuclear
shell model attributing each nucleon with the quantum numbers known for electrons. “Paired” nu-
cleons minimize overall spin according to wave function characteristics, and thereby improve bind-
ing energy.
3.3 The liquid drop model 75
In contrast, (odd, odd) combinations of (Z, N) are very rare, and “seem” to cre-
ate a problem for forming stable nuclei. In this case, nucleon binding energy is in-
terpreted as “negatively” affected. Unpaired nucleons (like the single “unpaired”
proton in Fig. 3.9) contribute to nucleon interactions in a different way and tend to
decrease nucleon binding energy. The coefficient thus has a negative sign.
If one sort of nucleon, i.e. either the protons or the neutrons, is of even number
while the other is of odd number, the “positive” and “negative” impacts cancel out
and the coefficient becomes δLDM = 0. The pairing term for those ca. 100 stable nu-
clei then simply disappears.
Fig. 3.9: Pairing term: two protons and two neutrons may
be considered as forming pairs inside a nucleus. Existing
virtually as pairs, this may optimize overall spin
quantum numbers (as introduced for electrons in the
Hund rule). Unpaired nucleons, like the “unpaired”
proton shown here, thus contribute to a weaker nucleon
binding energy.
The aim of the WEIZSÄCKER equation was to reproduce the experimental data of type
ĒB = f(A) shown in Fig. 3.4. Based on the LDM, the equation makes use of five (or
more) terms reflecting some intercorrelations found in the stable nuclei. The equa-
tion itself is a function of mass number and mathematically is a polynomial. The
terms of the WEIZSÄCKER equation are various multiples of individual powers of
A. The specific values of the individual coefficients of each term (αLDM, βLDM, γLDM,
ζLDM, and δLDM) are finally derived from a mathematical fit of the polynomial with
the (A, ĒB) dataset. Each of the coefficients carries the unit of energy (MeV) and a
corresponding sign. Table 3.7 summarizes the values of these coefficients.
Figure 3.10 shows the effect of including subsequent terms of the WEIZSÄCKER
equation, weighted by the coefficients. The impact of the pairing term is illustrated
for the “minus” effect, and had the “plus” effect been modeled, the final curve would
be above the yellow line. Likewise, for most of the unstable nuclides with the “zero”
effect, the curve would overlap with the yellow line.
Finally, Fig. 3.11 compares the mathematical model of mean nucleon binding
energy versus mass number A with the experimental data, indicated by individual
values of ĒB for the stable nuclei. Clearly, the LDM is able to reproduce the increase
76 3 Nucleons: Binding energy and shell structure
16
+ VOLUME TERM
14
- SURFACE TERM
12
10
- COULOMB TERM
ĒB (MeV)
8
- ASYMMETRY TERM
- PAIRING TERM
0
0 50 100 150 200 250
A
Fig. 3.10: Output of the WEIZSÄCKER equation when successively using the five basic terms (shown in
simplified and smoothed version).
in ĒB with increasing mass number A, a trend seen until ca. A = 50. It also captures
the relatively constant values of ĒB for mass number ranging from A about 50 to
about 100. Further, it nicely illustrates the impact of the last two terms, asymmetry
More terms: The classical WEIZSÄCKER equation included the five terms discussed. As nuclear
science research is an ongoing process, and the experimental database is improving (modern
atomic mass measurements provide more precise values of mass excess and nucleon binding en-
ergy), there are several modifications of the original version. This concerns both the value of the
coefficients and additional terms, such as “promiscuity”.
3.4 The nuclear shell model 77
and pairing, in order to reproduce the decrease in mean nucleon binding energy at
mass numbers of A > 100.
Table 3.8 compares the experimental values of ĒB with those calculated by eq.
(3.1) and gives the deviations ΔĒB between calculated and experimental values,
both in MeV and in percentage. For most isotopes of A > 20, there is excellent agree-
ment between experiment and theory. This is a mark of success for any theory and
indicates that rather classical physical relationships are working well.
However, some nuclei with specific mass numbers are not well fitted at all. This
is true in particular for the low mass number region (see Fig. 3.11 and the values
e.g. 4He in Tab. 3.8). Obviously, there is a limitation in the LDM’s ability to model
the mean nucleon binding energy.
10
6
ĒB (MeV)
0
0 50 100 150 200 250
A
Fig. 3.11: Mathematical model of the WEIZSÄCKER equation versus mass number A (red line), shown
alongside “real” mean nucleon binding energies ĒB (black squares). Experimental ĒB values are
shown for stable nuclei and the three uranium isotopes 234, 235 and 238 and 232Th.
Tab. 3.8: Experimental values of ĒB compared to those calculated via the liquid drop model and the
WEIZSÄCKER equation of type (3.1). Deviations ΔĒB between calculated and experimental values are
expressed both in MeV and in percent. In most cases, the liquid drop theory is predictive of mean
nucleon binding energy, i.e. fits well with experimental data.
ĒB ĒB ΔĒB
experimental calculated by eq. (.) calculated − experimental
He . . −. .%
C . . −. .%
O . . −. .%
Fe . . +. .%
U . . +. .%
went unexplained.23 Figure 3.12 shows the occurrence of stable nuclei with “magic”
numbers.
Because nuclei with one “magic” number (Zmagic or Nmagic) or two (double-magic,
magic
Z and Nmagic) show more stable isotopes compared to non magic nucleon compo-
sitions, there must be a reason related to nucleon binding energy. What may be re-
sponsible for such significant deviations in mean nucleon energy for some stable
nuclei? What do these nuclei with “magic” numbers have in common?
A statistical analysis of these nuclei shows that not only is there an unusually
high probability of finding the six “magic” numbers within the stable nuclei, but in
parallel, nuclei composed of protons or neutrons with Zmagic ±1 or Nmagic ±1, respec-
tively, are unusually rare. For illustration, Fig. 3.13 gives another excerpt of the Chart
of Nuclei. It covers the range around the “magic” number of Zmagic = Nmagic = 20. There
are six stable nuclei for Zmagic = 20 (the calcium isotopes, 40Ca, 42Ca, 43Ca, 44Ca, 46Ca,
48
Ca), but only three for Z = 19 (39K, 40K, 41K), and only one for Z = 21 (45Sc). There are
five stable nuclei for Nmagic = 20 (36S, 37Cl, 38Ar, 39K, 40Ca), but none for N = 19 and
only one for N = 21 (40K).
The LDM is not able to explain this “magic” number effect. Explanations arise
from a completely different point of view – the nuclear shell model (NSM). The two
most conclusive postulates are:
1. The nucleons are not distributed homogeneously, but in specific “shells”.
2. In contrast to the LDM, the NSM supposes that all the protons and all the neu-
trons are different from each other, i.e. they have individual characteristics that
make each nucleon in an atomic nucleus unique. This is the PAULI principle (in-
troduced for electrons) now applied to nucleons.
There is no “magic” number 126 for protons – such a chemical element does not exist.
3.4 The nuclear shell model 79
Z = 82
Z = 50
Z = 28
Z = 20
Z=8
N=8
N = 20
N = 28
N = 50
N = 82
N = 126
Fig. 3.12: Expression of stable nuclei with “magic” proton and neutron numbers indicated by
arrows. The “magic” numbers are 8, 20, 28, 50, 82 for Z and N and, in addition, 126 for N. Stable
nuclei of Z = 126 are not (yet) known. The N = 20 and Z = 20 section is enlarged in Fig. 3.13.
45Sc
100
35Cl 37Cl
75.76 24.24
31P
100
N=20
Fig. 3.13: Excerpt from the Chart of Nuclei covering stable nuclei around the “magic” number 20.
There are six stable nuclei for Zmagic = 20, but only three for Z = 19 and one for Z = 21. There are
five stable nuclei for Nmagic = 20, but none for N = 19 and only one for N = 21. (40K is not “really”
stable; its half-life is 1.28⋅109 years; comparable with 238U).
80 3 Nucleons: Binding energy and shell structure
Recall the development of the atomic model in general, and in particular the
change from the RUTHERFORD atom to the BOHR atom. It was quantum mechanics
that finally addressed this “individualism” of every electron by introducing shells
and shell occupancies derived from quantum numbers. The same theory must now
be applied to the protons and neutrons inside the nucleus.
For example, the experimental database says that nuclei having 2 or 8 protons
and/or neutrons appear to be particularly stable. Why not believe that 2 and 8 nu-
cleons each “fill” a “nucleon shell”? The equivalent would be electrons filling the K
shell with maximum 2 electrons (first period of PTE) and the L shell with 8 electrons
(second period of the PTE). This would be very interesting – if true!
Beyond a solely statistical distribution of numbers of stable nuclei in terms of
an excess along Zmagic and Nmagic, or a deficit in terms of Zmagic ±1 and Nmagic ±1,
there are several experimental facts that point to the special impact of the “magic”
numbers. In a sense, they extend the experimental database listed in Tab. 3.2 in the
context of the LDM and are listed in Tab. 3.9.
Tab. 3.9: “Experimental” observations that support a shell model of the atomic nucleus. Continued
from Tab. 3.2.
Observation
. Especially in the low mass region, there are several “peaks” along the fit of EB = f (A).
Obviously, some mass numbers create higher values of ĒB than predicted by the LDM (see
Fig. .).
. Nuclei with “magic” proton and neutron numbers (and the chemical element behind these
nuclei) are more abundant in the universe than nuclei/elements with proton or neutron
numbers, e.g. Zmagic ± and Nmagic ±.
. The nucleon separation energy of “magic” nuclei is extremely large (see Fig. .).
This is similar to the known excitation and ionization energies for electrons in the shell of an
atom (it is “difficult” to ionize a noble gas because its outer electron shell is full, and thus
energetically favored compared to an e- composition).
. Excitation energies of nuclei composed of “magic” nucleon number(s) are extremely large,
similar to the known excitation energies for electrons located in fully occupied shells or
subshells.
. Similarly, it is energetically inexpensive to add e.g. one neutron to a nucleus of Nmagic −
composition. This refers to a nuclear reaction process and is expressed as increased nuclear
reaction probability (“cross-section value”). In contrast, “neutron capture” probabilities are
low for Nmagic nuclei (see Chapter ).
3.4 The nuclear shell model 81
Argument no.11 of Tab. 3.9 concerns the separation energy of nucleons. This paral-
lels the discussion on ionization energies of chemical elements (see Chapter 1).
What is separation energy? While there is (in general) a gain in binding energy
when single nucleons enter the bound state of a nucleus, this is not an endless pro-
cess. Ultimately, the mixture of A, Z, and N becomes too unbalanced for one more
neutron or proton to be added. In a sense, the terms of the WEIZSÄCKER equation
(excluding the volume term) reduce the nucleon binding energy below some critical
level. Thus, this “last” nucleon is relatively easy to reject from the ensemble of ex-
isting nucleons. Separation energy describes the situation where the balance of en-
ergy is in favor of releasing a nucleon instead of keeping it. This is comparable to
the low first ionization energy of the alkali elements, and the high ionization ener-
gies of the noble gases (see Fig. 1.11).
The partial separation energy Es of nucleons is defined by eqs. (3.2) and (3.3).
Values of Es are high for a nucleus with a well-balanced mixture of protons and neu-
trons, and in particular for constellations of “magic” nucleon numbers. In contrast,
the separation energy is low in nuclei of one proton or neutron above a “magic” num-
ber. Figure 3.14 illustrates this pattern for nuclei around Nmagic = 8 and Zmagic = 8.
Argument no.13 of Tab. 3.9 may be considered as the reverse of nucleon separation,
namely the capture of one more nucleon. This is particularly relevant to the capture
of neutrons. Figure 3.15 illustrates this case for two of the stable strontium isotopes,
87
Sr and 88Sr. The two neutron capture reactions are 87Sr +1n→88Sr and 88Sr +1n→89Sr.
The nucleus 88Sr contains 50 neutrons, representing a “magic” number. When receiv-
ing another neutron, the “magic” number 50 turns into 51. Correspondingly, the “re-
action constant” of this nuclear reaction is low (0.005 b).24 In contrast, 87Sr may turn
its “non magic” neutron number of 49 into “magic” 50 when capturing one addi-
tional neutron. This reaction is much more probable than the other, and the reaction
constant is 16 b, i.e. larger by a factor of 3200.
See Chapter 13 for nuclear reaction constants expressed as cross-section values (σ), given in
units of barn (b). The cross-section σ is expressed in units of 10−28 m2, which is 1 barn.
82 3 Nucleons: Binding energy and shell structure
ES(p) 17F
0.600
15N
15.992
Fig. 3.14: Separation energies of nuclei around Nmagic = Zmagic = 8. Neutron separation energies
from 18O→17O→16O→15O are 8.044 MeV, 4.143 MeV, and 15.664 MeV, respectively. Proton
separation energies from 17N→16O→15N are 0.600 MeV and 12.127 MeV. An unusually high amount
of energy is needed to separate a proton or a neutron from the double-magic nucleus 16O:
15.664 MeV to remove one neutron, and 12.127 MeV to separate one proton. In contrast, (Nmagic +1)
and (Zmagic +1) nuclei such as 17O and 17F are eager to separate the one nucleon which is in excess
of a “magic” nucleon number. For example, the nucleus 17F will separate one proton given only
0.6 MeV.
N= 49 50 51
Fig. 3.15: The probability of neutron capture via a nuclear reaction process for two strontium
isotopes. The stable isotope 87Sr is of (Z, N) composition (38, 49), and the isotope 88Sr is of
(38, 50). N = 50 is a “magic” number. Both stable isotopes may accept a further neutron, which
creates a (Z, N +1) composition: 87Sr +1n = 88Sr, and 88Sr +1n = 89Sr. The cross-sections are
σ = 16⋅10−28 m2 and σ = 0.0058⋅10−28 m2, respectively. Obviously, 87Sr is much more likely to
capture a neutron than 88Sr. This is because 87Sr would reach a “magic” neutron number, while
88
Sr would lose its “magic” number.
3.4 The nuclear shell model 83
Obviously, the observations in Tab. 3.9 lead one to assign protons and neutrons to
specific shells within the nucleus according to a set of quantum numbers – similar
to electrons, populating specific orbits in the shell of an atom according to character-
istic quantum numbers. However, there should be some specific aspects for electrons
and nucleons, as, of course, the type and the energy of interactions are different for
electrons and nucleons. For nucleons, some features are listed in Tab. 3.10.
Main (n) The main quantum number n defines the shell that the protons or neutrons
exist in, depending on the energy of the shell.
Orbital (l) Nucleon shells are structured into subshells of orbital quantum number l,
and there are ≤l ≤(n −) subshells. Identical to electron orbitals, two
nucleons can fit in each s-shell. Likewise, there are , , , , , and
positions available for the l = p, d, f, g, h, and i shells, respectively.
However:
(ii) when naming nucleon orbitals, notation may follow the convention used
for electrons, or it may follow a different nomenclature system. In this new
system, the first occupied subshell l (according to increasing energy) gets
the number . This results in p and d notations – for an example, see
Fig. .(a) vs. (b).
Magnetic (m) The magnetic quantum number m reflects the projection of the orbital
angular momentum of the nucleons within a magnetic field. Possible values
are m = − l, . . ., l.
Spin (s) The spin quantum number of nucleons is either +½ or −½. Every energetic
and angular momentum state of an electron characterized by quantum numbers
n, l, and m thus does not accept more than two either protons or neutrons. Both
get the same set of (n, l, m), but differ in angular spin of +½ or −½.
While the HUND rule arranges electrons in subshells by first filling electrons
of the same spin into available orbitals (maximizing the total spin), nucleon
subshells are filled by initially pairing nucleons with opposite spin
(minimizing the total spin of the nucleus).
This was already seen for electrons in the p subshell, for example. Although of different orbital
quantum numbers (px, py, pz), their energies are identical.
84 3 Nucleons: Binding energy and shell structure
Total angular Interaction of orbital and angular momentum within the magnetic field of the
moments (j) nucleus results in a total angular quantum number j. This is comparable to
spin-orbit coupling. The possible values are j = l ±½. The total angular
momentum defines differences in energy. Levels initially defined by (nl)
quantum numbers form levels of ΔEls.
While this is (in principle) the same effect introduced for electrons, the
consequences are much more pronounced for nucleons. Fine structures
become more important.
With this set of quantum numbers, each nucleon is fully identified by charge
(+1 or 0) and energy, according to parameters described by quantum numbers (n,
l, m, +½) and (n, m, l, −½), or (j, +½) and (j, −½), as well as by s = ±½. Thus, PAULI’S
exclusion principle applies: in any quantum mechanical system (electrons in the
atom’s shells, nucleons in the atom’s nucleus), two particles with identical quan-
tum mechanical parameters should not exist.
Protons and neutrons occupy separate shells. This is because the two different nu-
cleons have different potential well considerations – see Fig. 3.16, which illustrates
the “particle in a box” model for nucleons, also called “infinite potential well” or
“infinite square well”. The well identifies, similar to the concept introduced for
electrons (Fig. 1.16), the distance over which a body exists at a local energy mini-
mum. The dimension of the well is very narrow, reflecting the size of the atom’s
nucleus. In contrast to a particle, which may “travel” unrestricted within a wide po-
tential well, nucleons may occupy certain positive energy levels only within a very
limited dimension. In classical physics, a particle is analyzed mathematically in
terms of its motion (or oscillation) when acted on by an external force. The parame-
ters are sinusoidal with constant amplitude and constant frequency. In quantum
mechanics, the particle is described as a wave function. The particle in a box model
can be solved analytically, depending on the model and its parameters, such as the
well width and the nucleon energies.
This situation considers nucleons already bound within the nucleus. A different
approach is needed to consider the moment a “free” nucleon enters the nucleus.
Here, the nucleon must approach a certain distance to the well, where it becomes
attracted by the strong force. This process is different for neutrons and protons due
to coulombic repulsion; an incoming proton is repelled by the positive charge of
3.4 The nuclear shell model 85
other protons in the nucleus (see Fig. 3.16). Finally, the potential energy of the nu-
cleons decreases when they enter the nucleus (the well). In this “bound” case, the
previous discussions on mass defect and nucleon binding energy apply.
Distance Distance
Potential well
Nucleus
with radius
r
Fig. 3.16: The particle-in-a-box model for a “free” nucleon entering the nucleus. The task is to fill
neutrons (white) and protons (orange) into shells and subshells of a nucleus in such a way that
shell occupancy mirrors “magic” nucleon numbers.
The potential well concept is rather abstract, particularly for wells with an infinite
square potential. To more accurately reproduce the potential energy of bound nu-
cleons, modifications must be made. One alternative to the “infinite square” poten-
tial is the “harmonic oscillator” potential. In addition, there is also a “realistic”
potential. Figure 3.17 schematically compares the three approaches.
The concept of the “infinite square” potential reproduces the observations indi-
cated in Tab. 3.9, where subshells are defined by “standard” quantum numbers,
leading every shell and subshell to have a specific (quantized) energy. Building
from this framework, the “harmonic oscillator” potential also allows specific energy
values to degenerate. Degeneration describes the situation when particles such as
nucleons (although of different quantum states) have the same energy. Discrete en-
ergies of the “infinite square potential” become equal. Quantum mechanics handles
this effect mathematically by the Hamiltonian for the particular system of more than
one linearly independent eigenstate with the same eigenvalue. As a result, the num-
ber of available orbitals is reduced, and nucleons should exist at lower, discrete en-
ergy levels with higher probability. Some initial quantum states subsume into one
energy level, which is filled by nucleons with equal probability. It no longer matters
which subshell the nucleon was in originally.
86 3 Nucleons: Binding energy and shell structure
U(r)
INFINITE SQUARE
HARMONIC
REALISTIC
Fig. 3.17: Variations on the potential well: infinite square potential, harmonic
oscillator potential, and “realistic” potential.
One may now start to puzzle the nucleons together. Let us see what happens when
the concept of quantum numbers is applied to nucleon shells, and a certain number
of protons and neutrons are put into the potential well. The aim is to see which
combinations of Z and N result in fully occupied shells and subshells, mirroring the
“magic” number isotopes (see Fig. 3.18 for the strategy).
Like electrons, nucleons are “filled” into the lowest-energy shells first. How-
ever, note that the arrangement of shells according to energy level is different, fol-
lowing: 1s ⇨ 2p ⇨ 3d ⇨ 2s etc. (as shown in column (a) of Fig. 3.19). The subshell
nomenclature used in nuclear sciences differs from quantum number descrip-
tions used for electrons. Every time a new (sub)shell appears, it is indicated by
consecutive numbers: 1s ⇨ 1p ⇨ 1d ⇨ 2s etc. (column (b)). The energy difference
between shell numbers decreases with increasing shell number.
3.4 The nuclear shell model 87
COMPLETELY FILLED
(OUTER) SHELL
Fig. 3.18: Strategy of the correlation of “magic” nucleon numbers with a nucleon shell model. A
completely filled (outer) nucleon shell should correspond to “magic” nucleon numbers (and vice
versa). A completely filled (outer) nucleon shell then represents higher stability of the nucleus.
This approach is exemplified for proton levels in Fig. 3.19.26 Two model poten-
tials are tested – the infinite square and the harmonic oscillator – resulting in dif-
ferent shell occupancies. The infinite square model considers the energy of all the
nucleons, as derived from their quantum numbers.
In contrast, the harmonic oscillator model allows for different sets of quantum
numbers to be identical in energy. It shifts the orbital energy levels known from the
infinite square approach. Second, nucleons within these “degenerate” levels have
identical energy. Finally, the number of energetically equivalent nucleons within a
common degenerate shell level is often larger than the shell occupancies of the
other model.
What matters here is the cumulative number of proton orbitals with a closed
shell. Similar to the concept that electron main shells are composed of individual
subshells, nucleon shell closures should be defined for shell levels separated by a
“sufficient” energy gap. For both models, these numbers are derived and given in
the black boxes of Fig. 3.19.
The model refers to spherically shaped nuclei which are also at ground state (i.e. not energeti-
cally excited).
88 3 Nucleons: Binding energy and shell structure
b a c d a c d
2g 18
3d 10
3p 6 138 3s 3d 2g 1i 2 10 18 26 168
1i 26
2f 14
3p 2f 1h 6 14 22 112
3s 2 92
1h 22
HARMONIC OSCILLATOR
INFINITE SQUARE WELL
2d 10
3s 2d 1g 2 10 18 70
1g 18 58
n=3
2p 6
2p 1f 6 14 40
1f 14 34
2s 1d 2 10 20
2s 2 20
1d 10
Energy
1p 6 8
1p 6 8
1s 2 2 1s 2 2
Fig. 3.19: Proton shells arranged according to the infinite square potential (left) and harmonic
oscillator potential (right) wells. The subshell arrangement used in nuclear sciences differs from
the ones used for electrons (a). For the sake of comparison, (b) indicates different energetic levels
for the n = 3 main shell. The theoretical maximum subshell occupancies are shown in white boxes
(c). Black boxes (d) show the resultant number of nucleons when the shells or subshells are filled
cumulatively. Obviously, both models are able to reproduce the “magic” numbers 2, 8, and 20 in
column (d). However, the shell closure numbers which follow (34, 58, 92, and 138 for the infinite
square well potential, and 40, 70, 112, and 168 for the harmonic oscillator model) are “non magic”.
Obviously, both models are able to reproduce the “magic” numbers 2, 8, and
20, as shown in column (e). However, the following numbers for completely filled
shells (shell closures) are 34, 58, 92, and 138 for the infinite square well potential,
and 40, 70, 112, and 168 for the harmonic oscillator. Both concepts thus fail to re-
produce the “magic” numbers >20.
Consequently, another approach is needed. Fortunately, the physics already ex-
ists: however, one must think beyond (sub)shell structures and (non)harmonic os-
cillation to the effect of fine structure, which relates to spin-orbit coupling (s-o-c).27
The interactions of orbital and angular momentum within the magnetic field of the
which was experimentally already known and described theoretically in the case of electron shell
structures and shell fine structures: the coupling of orbital and angular moments, see Fig. 1.21.
3.4 The nuclear shell model 89
nucleus results in a total angular momentum J. The total orbital quantum numbers
are j = l ±½. The total angular momentum defines differences in energy relative to the
initial, i.e. not spin-orbit coupled energy levels. Levels initially defined by (nl) quan-
tum numbers form new energetic levels of ΔEls. For example, the initial p-orbital (l = 1)
with altogether six protons (according to the three options of m = − 1, 0, and + 1, each
filled with two protons of s = +½ and −½) forms two new levels of j = 1 +½ = 3/2 and
j = 1 −½ = ½. These are filled with four and two protons, respectively (shell occupan-
cies after s-o-c are according to 2 j +1.) Likewise, the d-orbital splits into j = 5/2 (6 pro-
tons) and 3/2 (4 protons).
Spin-orbit coupling results in almost twice as many individual shells of charac-
teristic energy. All l-quantum numbers except l = 0 (the s-orbital) split into two lev-
els, i.e. double the number of shells. Moreover, their energy levels are rerranged.
Initially, the 1d-level was found above the 2s-shell, and protons had to be filled ac-
cordingly: first two protons in the 2s-shell, followed by 10 protons in the d-shell.
After spin-orbit coupling, the 1d-derived j = 5/2 level is filled first (with 6 protons),
followed by the 2s-shell (2 protons), and finally the j = 3/2 shell (4 protons). This
principle has the smallest effect on low energy level shell configurations. Differen-
ces in energy for the first shells are relatively large, at least larger than the effect of
splitting one initial level into two new ones. Thus, the effects of spin-orbit coupling
are operative only at higher shells.
In fact, spin-orbit coupling creates an extreme pattern of fine structure at
higher energy shells – exactly where they are needed to “correct” shell occupancy
numbers to better match the “magic” numbers. Consequently, the three cumulative
effects of new and energy shell open a great chance to “puzzle” again. Indeed,
Fig. 3.20 finally gives the optimum result: this model reproduces all the “magic”
numbers.
When comparing proton and neutron levels, there are two facts to note. First, the
energy levels of identical proton and neutron shells are slightly different. Second,
there are more shells available for neutrons to occupy – mainly because there are
more neutrons than protons in heavy nuclei. This corresponds to the increasing ex-
cess of neutrons over protons at mass numbers of A > 40, and is a constituent part of
the WEIZSÄCKER equation.
3.4.6 Energies of nucleon separation and excitation: from ground state to excited
state nuclei
Figure 3.14 illustrated the nucleon separation energy, ES, of stable nuclei around
Nmagic = Zmagic = 8. The individual values of ES ranged from 0.6 MeV to 15 MeV.
This is a representative range for the process of nucleon separation, which is
equivalent to atomic ionization for electrons (the latter are of keV energy only).
90 3 Nucleons: Binding energy and shell structure
a S-O-C d
11/2 12
1i
1/2 2
3p 3/2 4 126
13/2 14
5/2 6
2f
7/2 8
9/2 10
1h
3s 1/2 2
3/2 4 82
11/2 12
2d
SPIN-ORBIT COUPLING
5/2 6
7/2 8
1g
10
9/2
2p 1/2 2 50
5/2 6
3/2
4
1f
7/2 8 28
2s 3/2 4
2
20
1/2
1d 5/2 6
Energy
1/2 2
1p 3/2 4 8
1s 1/2 2 2
Fig. 3.20: Proton shells arranged according to spin-orbit coupling (s-o-c) relative to initial shell
occupancies (b). The white boxes show the maximum number of protons that can occupy an individual
shell/subshell after spin-orbit coupling. The black boxes (e) again show the resulting number of
nucleons when shells/subshells are filled cumulatively – these now represent the magic numbers!
3.5 Outlook 91
Similar to the processes of excitation known for electrons, nucleons may not
escape from the nucleus, but occupy higher energy levels as a consequence of exci-
tation.28 This important effect is typical for nuclei where Z or N > 20, as spin-orbit
coupling creates an increasingly dense overlap of upper subshell levels.
Nucleon excitation energies are thus lower than nucleon separation energy, but
again cover a wide range. Unlike electrons, excited nucleon levels may even persist
for relatively long periods. This results in differences between nuclei in the ground
state and excited states.
3.5 Outlook
The LDM treats protons and neutrons as identical particles, differing only by charge.
Based on this model, the semi-empirical WEIZSÄCKER equation used rather classical
ideas to describe nucleon interactions. Nevertheless, the calculated mean nucleon bind-
ing energies fit very well with experimental data – except for a few nuclei composed of
“magic” proton and neutron numbers. The NSM, in contrast, considers all nucleons to
be unique according to a set of quantum numbers. Once again, quantum mechanics
( in particular spin-orbit coupling) is key to modeling the “correct” arrangement of
shells in terms of energy and occupancy, reproducing all the “magic” numbers.
Suppose each field of a stable nuclide within the Chart of Nuclides represents a
black box, with only one face indicated. Here we get the parameters discussed so far.
However, both LDM and NSM provide much more information on each stable nuclide,
such as proton or neutron separation energies. Furthermore, for unstable nuclides,
much more information will become relevant – filling every surface of the virtual box.
These two nuclear models explain most properties of our material world at the
atomic scale. For example, it explains why we live in a world with an oxygen atmo-
sphere. Why? There are several stable isotopes of oxygen (16O, 17O, and 18O – see
Fig. 3.21). The most abundant is 16O, representing 99.762% of the mixture of the
three stable isotopes. The mean molecular weight of oxygen is 15.992 g/mol, as indi-
cated in the Periodic Table of Elements. This is because 16O is a double-magic nu-
cleus (8 protons, 8 neutrons). Its abundance on earth is much higher than expected
for non magic (Z, N) nuclei.
The fate of excited nucleons will become particularly important in Chapter 12, when discussing
post-effects of transformation (decay) processes of radioactive nuclides.
92 3 Nucleons: Binding energy and shell structure
7N 8O 9F O 16 O 17 O 18
14.01 15.999 18.998 99.762 0.00038 0.002
15 P 16S 17 Cl
30.97 32.07 35.45
Fig. 3.21: The Periodic Table of Elements and the Chart of Nuclides provide complementary
information on an element and its isotopes. As shown above, the molecular mass of oxygen
indicated in the PTE corresponds to the stable oxygen isotopes and their mixture.
The concepts discussed in this chapter are simplified, assuming that nuclei are
spherical and energetically non excited (ground state). Nucleons themselves are
considered to be solitary units, with no inner structure inducing specific behavior.
They are also assumed to have quasi-inert behavior and are not dependent on the
individual properties of neighboring nucleons. The mathematics describing nucle-
ons in this way is non relativistic.
Special binding energy for outer-shell (valence) nucleons, “Independent” particle model
particularly unpaired nucleons. This assumes that they are
different from the nucleons of inner shells, which form an “inert”
core
Interactions between all nucleons, including those in lower shells “Collective” model
and in pairs
Experimental evidence on deformation of nuclei and the effects on NSM of deformed nuclei
shell energy (“NILSSON” model)
3.5 Outlook 93
20 (3/2)+
(1/2)+
d3/2 (5/2)+
N=2 s1/2 (1/2)+
d5/2 (3/2)+
(1/2)+
8 _
(1/2)
_
p1/2 (3/2)
N=1
_
2 p3/2 (1/2)
Thus, theories on atomic nuclei are under continuous revision. Consequently, there
are many ongoing discussions on alternatives to the LDM and NSM. Important exten-
sions are listed in Tab. 3.11. In general, for a deeper understanding of the properties
and behavior of a nucleus, the composites of the nucleons (elementary particles)
must be considered, as well as the forces that allow subatomic particles to interact.
These interactions are needed to describe processes of transformations between dif-
ferent states of an atomic nucleus and the atom in general. Primary and secondary
94 3 Nucleons: Binding energy and shell structure
Aim: Until now, more than 3200 nuclide species have been identified, and less than
10% of them belong to the class of stable nuclides. Therefore, approximately 3000
known nuclides are unstable. They show compositions of proton and neutron num-
bers that do not provide mean nucleon binding energies sufficient to keep the en-
semble of nucleons stable forever. Instead, unstable nuclides try to convert into
stable ones by either changing the ratio between protons and neutrons or the abso-
lute number of nucleons.
This transformation is always driven by the gain in ΔĒB between an initial (K1)
and a formed (K2) nucleus. Processes of transformation of an unstable nucleus into
a stable (or at least a “more stable”) one are always exothermic. Transformation
processes start with primary processes (changing nucleon compositions). In many
cases, primary processes populate excited nuclear states of K2, which either directly
or with some “delay” further stabilize via secondary processes. Both primary and
secondary transformations may induce follow-up (post) processes.
Actually, the primary, secondary and post-transformation processes are the
source of what is called “radioactivity” and unstable nuclei are the “radioactive”
ones.
This chapter aims to introduce a kind of systematics of/between the three main
primary transformations. The following chapters then will focus on the individual
processes in detail.
The strong interaction responsible for nucleon binding energy is the driving force
behind nuclide stability. This nucleon binding energy correlates with numbers of A,
Z, N of a nucleus (theoretically well described by the LDM) and the internal struc-
ture of its nucleons (covered by the NSM) – based on the data available for stable
nuclides. Figure 3.11 for example shows that for each mass number A there is a cor-
responding value of ĒB representing stable nuclei. Nuclei with these values of ĒB(A)
show a certain mixture of protons and neutrons – guaranteeing stability. So what if
the mixture deviates from that level? The WEIZSÄCKER equation will yield values of
ĒB less than the corresponding maximum value ĒB(A). Consequently, such nuclei
will not be stable.
Let us see how changes in nucleon composition develop and finally leave stable
configurations. Figure 4.1 gives an excerpt of the Chart of Nuclides for the segment
https://doi.org/10.1515/9783110742725-004
96 4 From stable to unstable nuclides
Z=
11
10
N= 4 5 6 7 8 9 10 11 12 13 14 15 16
Fig. 4.1: Segment of the Chart of Nuclides. Stable nuclides are marked in black. White fields
indicate “vacancies” available for isotopes of corresponding (N, Z) configuration. Lower numbers
are the relative abundances of the stable isotopes. The filling of these vacancies is illustrated for
isotopes of oxygen. Theoretically, there may be isotopes of oxygen of either 0 < N < 8 or N > 10. They
are filled with different intensities of gray. The closer the isotopes are to the stable nuclides, i.e.
15
O close to stable 16O, and 19O close to stable 18O, the darker the gray. The further the isotopes
are from the stable core, the lighter the gray. This decreasing color intensity follows the decreasing
mean nucleon binding energy and schematically corresponds to the decreasing stability of the
nuclides.
Figure 4.1 indicates that there may be isotopes of oxygen of very different neutron
number. Theoretically, this could be of either less than 8 (8 > N > 0), which is “neutron
deficit”, or more than 10, which is “neutron excess”. However, not all the theoretical
(N, Z = 8) combinations are feasible. Qualitatively, this is obvious. It would be an
4.1 Unstable nuclei and the Chart of Nuclides 97
extreme situation to imagine an oxygen nucleus composed of its 8 protons, but with-
out any neutron! Or just one or two neutrons! One would already “feel” that this
could never create a nucleus that could exist even for a short period of time. A semi-
quantitative explanation comes from the WEIZSÄCKER equation, in particular from the
coulomb term.
To quantify the feasibility, the concept is best illustrated for the process of add-
ing more and more neutrons to the 8 protons of the nucleus of oxygen. Yes, each
individual neutron added increases the overall nucleon binding energy partially
(δĒB), defined by the volume term of the WEIZSÄCKER equation: K1N−1 ! K2N.
However, the (N – Z)2 part of the WEIZSÄCKER equation increases with every fur-
ther neutron – and thus reduces the value of the overall nucleon binding energy of
the nucleus. Such nuclei may prefer to eliminate an “excess” neutron instead. This
is the neutron separation energy ES(n) needed to eliminate one neutron out of the
nucleus: K1N ! K2N−1. The energy is calculated by eq. (4.1) using absolute masses of
the two nuclei and the neutron.
Conceptually, there will be a situation in which the partial nucleon binding energy
gained by capturing another neutron through the volume term (δĒB) is the same or
even less than neutron separation energy ES(n), eq. (4.2).
This correlation allows one to discriminate between isotopes that theoretically
may exist and those that are simply not allowed to exist. For oxygen isotopes,
Fig. 4.1 graphically indicates the general and simplified1 tendency of nucleon bind-
ing energy among the isotopes of neutron excess and deficit. The same can be done
for neutron and proton separation energies ES(n) and ES(p).2 For the isotopes of ox-
ygen and carbon they are illustrated in Fig. 4.2.
There is a general conclusion: the more “in excess” one type of nucleon is to
the other, the easier it is to separate. And vice versa: the stronger the deficit of one
sort of nucleon compared to the other, the more energy is needed to separate it.
For oxygen, for example, the regions in which isotopes simply cannot exist at
all are at mass numbers A(8O) ≤ 12 and A(8O) ≥ 25. The crosses at the left and right in
Fig. 4.3 thus indicate the impossibility of the existence of these isotopes.
There is an alternating effect of nucleon binding energies for odd and even isotopes, discussed in
Chapter 3. Oxygen isotopes with an even neutron number are superior in mean nucleon binding
energy compared to isotopes with an odd neutron number.
These values can be calculated according to eq. (4.2), and are tabulated in nuclear science
literature.
98 4 From stable to unstable nuclides
35 35
30 30
25 25 Es (n) Es (p)
Es (n) Es (p)
20 20
Es (MeV)
Es (MeV)
15 15
10 10
5 5
0 0
-5 -5
6 8 10 12 14 16 18 20 22 24 10 12 14 16 18 20 22 24 26 28 30
Carbon isotope mass number A Oxygen isotope mass number A
Fig. 4.2: Neutron and proton separation energies ES(n) and ES(p) for the isotopes of carbon and
oxygen.
Consequently, the number of unstable isotopes allowed to exist is less than the
theoretically possible ones.3 Figure 4.3 shows the resulting isotopes of oxygen and
carbon. For oxygen, it indicates that among the light (neutron-poor) isotopes, four
(N, Z) mixtures remain, i.e. at 4 ≤ N < 8. For the heavier (neutron-rich) isotopes of
oxygen, the last one representing ES(n) < −δĒB is 28O, i.e. N = 22. Altogether, there are
(only) 4 “light” and 10 “heavy” unstable isotopes of oxygen. For carbon, there are 4
and 9, respectively. (The reason for generally more unstable isotopes in the case of
neutron excess compared to deficiency can also be explained by the LDM and NSM.)
The same principle applies to protons, i.e. the processes of K1N−1 ! K2N vs.
K1N ! K2N−1 along an isotone line. The concept, nevertheless, is better illustrated
by following an isotone line on the Chart of Nuclides. Another option is to discuss
the isobar line of the Chart of Nuclides. Figure 4.4 shows the same segment of the
Chart of Nuclides as in Fig. 4.3. This time, the isobar nuclides of mass number A = 16
are highlighted. It demonstrates that only 6 nuclides may exist along an isobar line.
Close to the only one stable nucleus 16O (Z = 8) at A = 16 (ĒB = 7.976 MeV), there are
two at larger Z (Z = 9 = 16F and Z = 10 = 16Ne), and three with lower proton number
(Z = 7 = 16N, Z = 6 = 16C, Z = 5 = 16B). Further combinations of Z:N, such as 16Na and
16
Be, are energetically impossible.
When the crosses in Fig. 4.3 are connected, two lines will appear: the one at
neutron excess is called the “neutron drip line”, the other one the “proton drip
Furthermore, the number of unstable isotopes experimentally identified is even smaller. Re-
search towards new, i.e. not yet detected nuclei is ongoing.
12O 13O 14O 15O 16O 17O 18O 19O 20O 21O 22O 23O 24O 25O 26O 27O 28O
99.762 0.00038 0.002
4.879 5.812 7.052 7.464 7.976 7.751 7.767 7.556 7.569 7.389 7.365 7.164 7.016 6.723 6.457 6.175 5.925
∆N = 4 ∆N = 10
∆N = 4 ∆N = 9
8C 9C 10C 11C 12C 13C 14C 15C 16C 17C 18C 19C 20C 21C 22C
98.90 1.10
3.098 4.337 6.032 6.676 7.680 7.470 7.520 7.100 6.922 6.558 6.426 6.118 5.959 5.659 5.959
Fig. 4.3: Unstable isotopes of oxygen and carbon theoretically allowed to exist, according to the balance of protons:neutrons, partial nucleon
binding energy and neutron separation energy – δĒB vs. ES(n). ĒB values are indicated in MeV. The crosses at the left and right indicate that these
isotopes cannot exist at all. When connecting the crosses at the neutron-poor branch and the neutron-rich branch, two lines will appear: The one at
neutron excess is called the “neutron drip line”, the other one the “proton drip line”.
4.1 Unstable nuclei and the Chart of Nuclides
99
100 4 From stable to unstable nuclides
16Ne
6.083
16F
6.964
12O 13O 14O 15O 16O 17O 18O 19O 20O 21O
7.976
16 N
7.374
10C 11C 12C 13C 14C 15C 16C 17C 18C 19C
6.922
16B
5.509
16Be
4.270
Fig. 4.4: Isobar nuclides of mass number A = 16. ĒB values are indicated in MeV. Only seven
nuclides may exist. Close to the only stable nucleus 16O (Z = 8) at A = 16 (ĒB = 7.976 MeV), there are
two at larger Z (Z = 9 = 16F, and Z = 10 = 16Ne), and three with lower proton number: Z = 7 = 16N,
Z = 6 = 16C, Z = 5 = 16B. Further combinations of Z:N, such as 16Na and 16Be, are energetically
impossible (red crosses).
line”. Figure 4.5 shows the two resulting areas of the Chart of Nuclide. The intense
gray area indicates the experimentally identified unstable nuclides, and the larger,
light gray area the theoretically possible ones.
When discussing the huge number of unstable nucleon compositions that can theo-
retically exist (relative to the number of stable ones), one may separate these unstable
nuclides into two classes. One class has a more “natural” character and includes
4.2 Unstable nuclides on earth 101
82
50
NEUTRON
28 DRIP LINE
20
8
2
2 8 20 28 50 82 126 N
Fig. 4.5: The Chart of Nuclides indicating nucleon drip lines of ES ≈ 0. The dotted line limiting the
lower area at neutron excess is the “neutron drip line”, the upper one the “proton drip line”. The
(red) area along the stable nuclides (solid lines) indicates the area of experimentally identified
unstable nuclides.
unstable nuclides that exist on our earth without human help, while the other class
covers nuclides that are created “artificially” by man-made nuclear reactions.
The first class of unstable nuclides includes a relative small number that are of
practical relevance in many fields (scientifically, analytically, industrially, medi-
cally). Although of natural origin, there are three very different subtypes according
to the process of origin and the “age” of the unstable nuclides.
The first sort includes a subgroup of nuclides of ancient age – the origins of
these unstable nuclides goes back to the formation of our planet and beyond. The
nuclides thus are of half-life of at least (or close to) the age of the earth (4.6⋅109
years). This group includes light nuclides such as 40K (t½ = 1.28⋅109 a), but mainly
very heavy nuclides such as isotopes of thorium and uranium, namely 232Th (t½ =
1.405⋅1010 a), 235U (t½ = 7.038⋅1008 a), 238U (t½ = 4.47⋅109 a). These nuclides trans-
form into others in order to stabilize (see Chapter 5 for “Naturally occurring trans-
formation chains”). They permanently and naturally generate a number of unstable
nuclides until they finally reach stable nucleon configuration at isotopes of lead or
bismuth. This is another subgroup comprising the successors of the three men-
tioned nuclides 232Th, 235U and 238U (sometimes also called “daughters” of the three
102 4 From stable to unstable nuclides
The other sort of naturally occurring unstable nuclides were not incorporated into
the earth’s crust in the course of the formation of the planet, but are being continu-
ously produced by cosmic radiation. Their origin is thus not in the earth’s crust, but
in its atmosphere. Cosmic radiation comprises a class of particle radiation and elec-
tromagnetic radiation, either emitted by the sun or arriving from the universe, that
reacts with the chemical elements constituting our atmosphere.4 The most prominent
example is the permanent formation of an unstable isotope of carbon, namely 14C,
within the nuclear reaction 14N(n, p)14C as induced by “cosmic” neutrons.5
What may be concluded from this sort of unstable nuclides is: unstable (i.e., ra-
dioactive) nuclides, transforming in the context of “radioactivity”, are a natural part
of our ecological system. Furthermore, radionuclides such as 40K and 14C co-exist
among stable isotopes of potassium (39K and 41K) and carbon (12C and 13C), respec-
tively, and are incorporated in our own biological matrix! A characteristic (although
mass-wise low) amount of our biological matter (for example for carbon: amino
acids, proteins, glucose etc.) is “radioactive”. So – this is what we are in part our-
selves: radioactive by nature.
None of the chemical elements beyond lead offer stable isotopes, so all of them are
radioactive. A few of them are generated in a natural way as successors of the fami-
lies of the long-lived nuclides of 232Th, 235U and 238U. All the isotopes of the chemi-
cal elements beyond uranium have been man-made.
The same is true for all the radioactive isotopes of proton number 82 (Pb) > A >
90 (Th) not involved in the natural transformation chains; and almost all of the un-
stable isotopes of the chemical elements ranging from hydrogen to lead.
The processes are quite complex and are in part covered by the Chapter 13 on “Nuclear reactions”,
but mainly in Chapter 2 on “Radiation dosimetry” and Chapter 3 “Nuclear dating” of Volume II.
The character of nuclear reactions is described in Chapter 13, and some relevant features of 14C
are covered in Volume II in the context of “nuclear dating”, Chapter 3.
4.2 Unstable nuclides on earth 103
Fig. 4.6: Radioelements and selected chemical elements co-existing with unstable isotopes.
Red: Radioelements are those elements existing on earth only in the form of unstable isotopes when
produced by artificial, man-made nuclear reactions, namely Tc, Pm, Pa and all the transuranium
elements. Green: Elements such as U, Th, Ac, Ra, Fr, Rn, At and Po exist on earth exclusively in the form
of more or less unstable isotopes. These are either permanently generated from long-lived nuclides
of uranium and thorium and thus are of endogenous origin, or may also be produced artificially.
Blue: Carbon has stable isotopes (12C and 13C), but co-exists with an unstable isotope (14C), which is,
firstly, permanently being generated by cosmic radiation in the earth’s atmosphere and, secondly,
incorporated into biological material. Hydrogen with its unstable isotope tritium (3H) is another example.
Brown: Potassium belongs to the “normal” chemical elements and was part of the earth’s formation.
However, one isotope (40K) is in fact unstable (although of extremely long half-life), is radioactive and
thereby emits radiation – but nevertheless constitutes an endogenous part of biological matter.
This may be no of great concern for “conventional” chemistry, as almost all the
chemical elements ranging from hydrogen to lead offer stable isotopes – sufficient to
investigate chemical properties and to cater to practical applications. However, there
are a few exceptions. Some of the chemical elements listed in the PTE do not have sta-
ble nucleon configurations and do not exist at realistic amounts in nature.6 Chemical
This explains – at least in part – why the original Periodic Tables of MENDELEVEV and MEYER (es-
tablished on the basis of impressive systematic experimental evidence) revealed some “empty
spaces” or “blank spots”.
104 4 From stable to unstable nuclides
4.3 Outlook
Various combinations of protons and neutrons create a little universe of stable and
unstable nuclides. The number of unstable nuclides, however, is limited due to the
balance of partial nucleon binding and separation energies. So far, about 3000 un-
stable nuclides have been identified compared to less than 300 stable ones.
However, all the unstable nuclides are in a continuous process of transforma-
tion towards one final destiny: to achieve a combination of proton and neutron
numbers guaranteeing “stability”. The way a nuclide “organizes” its transformation
is very complex. Usually, these types of transformation are also called “decay”,7
but nothing actually “decays”: it is all about transformations! Actually, the transfor-
mation itself is the source of what is called “radioactivity” and the unstable nu-
clides involved are the “radioactive” ones. All these unstable nuclides thus are also
referred to as “radionuclides”. The effects that accompany this transformation, fun-
damentally particle or electromagnetic emission, are called “radiation”.
As in nature in general, physical and chemical objects tend to minimize their
overall energy. This is also true for nucleon binding energy – and thus for nucleon
configurations of a nucleus of an atom. According to the WEIZSÄCKER equation, mean
nucleon binding energy changes whenever compositions of (A, Z, N) are modified.
This transformation is driven by the gain in ĒB between an initial (✶K1) and a formed
(K2) nucleus: ĒB(✶K1) = (Z, N) ! ĒB(K2) = (´Z, ´N´). It proceeds spontaneously and is ir-
reversible! The transformation processes are exothermic, as the newly formed nuclide
K2 has a smaller overall nucleon binding energy than the initial one. A more stable
nuclide will never spontaneously convert into a less stable descendant.
The various versions to realize these processes require separate and detailed
description in the following chapters. The different classes of transformation of one
unstable nuclide into another, more stable one and the kind of emissions that ac-
company the nuclear processes, nevertheless, all obey the same mathematical con-
cepts. These are introduced in the next chapter.
This textbook prefers the wording “transformation” to decay, although both expressions are
equivalent. The basic feature of unstable nuclei is to proceed into more stable states.
5 From stable to unstable nuclides: Mathematics
Let us consider a number Ń of one sort of an unstable nuclide ✶K1 which transforms
into a new nuclide K2: ✶K1UNSTABLE ! K2. This number will decrease because a new
nuclide is formed. A fraction of nuclides (dŃ) will transform in a fraction of time
(dt), the sign of which is negative: − dŃ/dt. The value of dŃ/dt represents energetic
https://doi.org/10.1515/9783110742725-005
106 5 From stable to unstable nuclides: Mathematics
parameters of the unstable nuclide, mainly the amount by which the mean nucleon
binding energy ΔĒB can be improved when transforming into a more stable nuclide.
For each nuclide, there is a probability for how fast the process is – the kinetics of
the transformation are specific to each sort of unstable nuclide.
Now it is possible to consider the fraction of unstable nuclides that have transformed
over a period of time, relative to the initial number of nuclides (at t = 0): − dŃ/Ń = f(t). If
the number of nuclides transformed is proportional to the period of time, dŃ/Ń ≈ dt.
Next, a proportionality factor λ is introduced: dŃ/Ń = −λdt. In order to guarantee con-
stant units on both sides of the equation, the unit of λ should by 1/t, a reciprocal of time.
This is typically expressed in terms of inverse seconds, i.e. s−1.
When integrating eq. (5.1), a quantitative correlation between the number of
unstable nuclides (Ń) still existing after any period of t and Ńo, the number existing
initially (at t = 0), can be calculated. Actually, the parameter “time t” in eq. (5.3)
should read as (t – to) or Δt, with Ńo identifying the number of nuclides at time to.
However, to may be taken as t = 0. In analogy to monomolecular reaction kinetics in
chemistry, this is a mononuclear kinetics of transformation of unstable nuclides:
nuclear transformations follow a first-order exponential rate law, see Figure 5.1.
dŃ dŃ
(5.2)
(5.1)
– = f(t) = dt
Ń Ń
(5.3)
Ń = Ńo e- t
Fig. 5.1: Basic consideration to derive the mononuclear transformation kinetics of unstable
nuclides: A first-order exponential rate law. Ń is the number of unstable (radioactive) nuclides
existing at any time t, relative to the number of unstable nuclides which had existed initially at
time t = 0. The proportionality factor λ is the transformation (or “decay”) constant.
It may be confusing to see that the symbols N and A are getting ambiguous: N in the earlier chap-
ters symbolized “neutron number”, now also “number of nuclei”. A was introduced as “mass num-
ber”, and now it is used for “activity” or “radioactivity” too. This is the reason why Ń and Á are
used instead in this textbook for radioactivity and number of transforming nuclides.
5.1 Transformation parameters 107
s−1. In parallel with the decreasing number of unstable nuclides over time, radioac-
tivity also decreases exponentially with e−λt, as per eq. (5.6), see Figure 5.2.
dŃ ^
− = Á (5:4)
dt
Á =λ Ń (5:5)
Á
λ= (5:6)
Ń
dŃ ‸
– =Á Á = Ń = Á/Ń
dt
Fig. 5.2: Radioactivity Á is derived from the kinetics of transformation of unstable nuclides into
more stable ones. Ń and Á are directly proportional. The proportionality constant is λ. The
corresponding unit of radioactivity is 1/time = s−1.
The decay (or transformation) constant λ has units of reciprocal time (s−1).2 The
transformation constant λ lends its unit to that of activity, shown in eq. (5.5). The
IUPAC system adopts this unit in the context of radioactive transformation, defining
it as 1 Bq (Becquerel) = 1 s−1. The value of 1 Bq signifies exactly one event of sponta-
neous radioactive nuclear transformation within 1 second. Because in many areas
of research or daily life the scaling of radioactivity is much higher, it is common to
use units such as MBq or GBq (mega-Becquerel or giga-Becquerel, respectively) –
see also Fig. 5.8.
5.1.3 Half-life
There are two other options to quantify the velocity of (radioactive) transformation.
One is “mean lifetime” τ as sometimes used in nuclear physics.3 This other one is
An identical unit is known from another area of physics to quantify “frequency” with the unit Hz
(Hertz), i.e. the number of oscillations (or other cycles) per second. For example, 1000 oscillations
per seconds = 1000 Hz = 1000 s−1. This describes a constant process, i.e. 1000 oscillations per sec-
onds = 1000 Hz = 1000 s−1. In contrast, kinetics of radioactivity follows a first-order exponential
law, so the meaning is different.
Nuclear physics may use an expression well-established in general mathematics, called “expo-
nential time constant”, “mean lifetime” or simply “lifetime”. It quantifies the period of time t,
108 5 From stable to unstable nuclides: Mathematics
“half-life”, which is much more common in nuclear and radiochemistry. The half-
life identifies the period of time t during which exactly half of the initial unstable
nuclides have transformed, as shown in eq. (5.7).4
Ń 1
= (5:7)
Ń0 2
within which an initial number of nuclides Ńo (or any other species) had transformed down to a
fraction of the reciprocal value of e (1/e = 0.367879411), i.e. Ń = 1/e Ńo. Similar to the use of “trans-
formation constant”, the corresponding exponential equation is Ń(t) = Ńo e−t/τ. The correlations are
τ = 1/λ = t½ / ln(2), and t½ = τ ln(2) = 0.693147181 τ. Consequently, all three parameters are in-
terchangeable. In the case of the carbon-14 isotope, for example, the transformation constant is
λ = 5.775⋅10−4 s−1, the half-life is t½ = 5730 years, and the mean lifetime is τ = 8267 years. However,
the use of “lifetime” is not applied to radio- and nuclear chemistry, and may be limited to elemen-
tary particles. For example, for free neutrons the mean lifetime is usually given as τ = 881.5 ± 1.5 s
(= 14.69 min), which is t½ = 611 ± 1.0 s (= 10.18 min).
What, for example, is described by a value of t½ = 1 h? Suppose there are Ńo unstable nuclides at
any time-point to (which we measure according to their radioactivity Áo). Exactly one hour later,
the number of unstable nuclides still in existance is exactly half: Ń(t +1 h) = Ńo/2 = 0.5 Ńo. The same
is true of the radioactivity “belonging” to these nuclides: Á(t +1 h) = Áo/2 = 0.5 Áo.
5.1 Transformation parameters 109
1
(5.3)
Ń = Ńo e- t
Ń
= e- t
Ńo
2
Ń 1
= = e- t½
Ńo 2
ln2 0.693 3
(5.9)
t½ = =
(5.8)
0.5 = e- t½
Fig. 5.3: Conversion of transformation constant λ into half-life t½. At t = t1/2 there is Ń = Ńo/2.
Ń = Ńoe−λt transforms (2) into a modified version of eq. (5.3). This proceeds via (2) Ń/Ńo = ½ = 0.5,
and substitutes time t by half-life t½. Via (3) and (4) the transformation constant λ and half-life t½
correlate.
half-lives steadily decrease by a factor of 2. After one half-life, exactly 50% of Ńo (or
Áo) remains. After another half-life, this is 25%. The values of Ńo (or Áo) are halved
after each additional half-life. Following 10 half-lives, the nuclide population is less
than 0.1% (0.09765625%) of its initial value. Mathematically, the fraction or percent-
age of the number of initial nuclides remaining after n half-lives is 1/2n or 100/2n,
respectively.
Exponential correlations can be converted from a linear scale to a logarithmic
scale, yielding instead of eq. (5.3) the linear relationships of type log(Ń) = log(Ńo) – λt,
as per eq. (5.10). This is true for the number of unstable nuclides, as well as their abso-
lute radioactivity Á and the (relative) count rate Ć of measured radioactive events (see
below).
Figure 5.5 identifies two elementary features. The linear function crosses the y-axis
at log(Ńo), indicative of the number of unstable nuclides Ńo at time-point 0 (or may
be extrapolated to any earlier time-point.) The line’s slope is negative and repre-
sents the value of λ. Thus, experimentally obtained trends like this quantify both
the half-life of a single unstable nuclide and its absolute numbers.
Among the unstable nuclides detected so far, there is an impressive range of
half-lives (or transformation constants). There are unstable nuclides of half-lives >109
110 5 From stable to unstable nuclides: Mathematics
100
0.1953125 %
0.097656 %
0.390625 %
90
0.78125 %
1.5625 %
3.125 %
12.5 %
6.25 %
50 %
25 %
80
70
60
Ń (%)
50
40
30
20
10
0
0 1 2 3 4 5 6 7 8 9 10
t/t½
Fig. 5.4: General trend in transformation of unstable radioactive nuclides vs. half-lives: Exponential
law of the kinetics of radioactive transformation. The x-axis representing time is on a linear scale
and is divided by half-life. The y-axis is on a linear scale and represents Ń. If the initial number Ńo
or the corresponding radioactivity Áo of a sort of unstable nuclide in existence at an initial time-
point of t = 0 is considered, the fraction of nuclides (or activity) still present after subsequent half-
lives steadily decrease by a factor of 2. After one half-life, exactly 50% of Ńo (or Áo) remains. After
another half-life, this is 25% etc.
years, which preceeds the existence of our planet. Extremely short half-lives ap-
proach the dimensions of fractions of a second. Table 5.1 collects some selected un-
stable nuclides, which are relevant for specific reasons,5 together with their half-lives
and transformation constants.
Individual radionuclides will be discussed in detail in the various chapters of this Volume and in
Volume II of this textbook.
5.2 Correlations between radioactivity, number of unstable nuclides and masses 111
gŃ
logŃo
o
1 x t½
-
logŃ
o
0 1 2 3 4 5 6 7 8 9 10
t/t½
Fig. 5.5: Linearized version of the first-order rate law of radioactive transformation kinetics. The
x-axis representing time is (still) on a linear scale (as in Fig. 5.4) and gives multiples of half-lives.
The y-axis is plotted on a logarithmic scale as log(Ń) instead of Ń. According to log(Ń) = log(Ńo) – λt,
there is a linear line with a negative slope, and its absolute value is λ. The intercept of the linear line
at y-axis gives log(Ńo). Instead of log(Ń), the y-axis could represent log(Á) or log(Ć).
Tab. 5.1: Selected unstable nuclides representing a broad range of half-lives and typical areas of
use in nuclear science research and applications.
Á M
Fig. 5.6: Correlations between radioactivity, number of unstable nuclides (and via the molar
mass M and mass m of the nuclides), and transformation constant.
5.2.1 From Ń to Á
Table 5.2 illustrates this situation for a fixed number of 106 unstable nuclides, but
various half-lives ranging from 1 s to 1 min, 1 h, 1 day and 1 year. Whereas in the case
of a short half-life (e.g. 1 min) the radioactivity is high (1.155⋅104 Bq), a long half-life
(e.g. 1 h) gives a radioactivity of almost 2 orders of magnitude less (1.925⋅102 Bq).
5.2.2 From Á to Ń
Vice versa, the same level of radioactivity may be created by a very different number
of radioactive nuclides. Table 5.3 compares the corresponding values for two different
radionuclides of carbon, namely the isotopes carbon-11 and carbon-14. Both differ sig-
nificantly in terms of λ and t½. The values of λ are 5.78⋅10−4 s−1 and 3.85⋅10−12 s−1,
5.2 Correlations between radioactivity, number of unstable nuclides and masses 113
t½ λ (s−) Á (Bq)
−
second .⋅ .⋅+
minute .⋅− .⋅+
hour .⋅− .⋅+
day .⋅− .
year .⋅− .⋅−
Tab. 5.3: Half-lives t½ and transformation constants λ of two different unstable, radioactive
isotopes of the chemical element carbon, namely carbon-11 and carbon-14. Both differ significantly
in terms of λ and t½.
Á = 1⋅106 Bq = 1 MBq
6.022⋅1023 Ñ = 1 mol = x g
isotope t½ λ s−1 Ń mol g
C min .⋅− .⋅ .⋅− .⋅−
C a .⋅− .⋅ .⋅− .⋅−
Following Á = λŃ, the number of radioactive nuclides for the same Á depends on
the value of the transformation constant λ. For larger λ values (or smaller t½ val-
ues, according to t½ ≈ 1/λ), fewer transforming nuclides are necessary to create the
same radioactivity in terms of transformations per second. Figure 5.7 illustrates the
One may say that 11C is a “short-lived” isotope, and 14C is a “long-lived” isotope. However, as
half-lives of the many unstable nuclei range from a fraction of a millisecond to billions of years –
everything is relative.
114 5 From stable to unstable nuclides: Mathematics
correlation of the half-life of a radioactive isotope and the corresponding mass in the
case of a constant radioactivity of 106 Bq (1 MBq). The half-lives listed in Tab. 5.2 are
shown separately and also the two longer half-lives of 100 and 109 years. The x-axis
gives values of t½ ranging from 100 to 1018 s, i.e. over 18 orders of magnitude! The
y-axis gives the corresponding number Ń of radioactive nuclides according to half-life.
It ranges from 106 to 1021 nuclides. Again, showing 15 orders of magnitude requires a
logarithmic scale. The red line is the correlation of t½ vs. Ń for the level of Á = 1 MBq.
1021 1021
1020 1020
1019 1019
1018 1018
1017 1017
1016 1016
Ń 1014 1014
1 year
1013 1013
1012 1012
1011 1011
1 day
1010 1010
1 hour
1009 1009
1007 1007
1 second
1006 1006
Fig. 5.7: Correlation between half-life t½ and number Ń of unstable nuclides for a constant
radioactivity. Half-lives of 1 second, 1 min, 1 h, 1 day, 1 year, 100 years and 109 years are shown
separately. Both the x-axis and y-axis are logarithmic. The line correlates t½ with N for the level of
Á = 1 MBq. (Other levels of radioactivities would create lines parallel to the one shown.) The square
indicates the range of half-lives typically handled in research institutions, i.e. from minutes to
years. Consequently, typical numbers of radioactive nuclides (for the Á = 1 MBq level selected)
range from 107 to 1016, i.e. many orders of magnitude less than, e.g., the 1 mole level of
6.022⋅1023.
5.2 Correlations between radioactivity, number of unstable nuclides and masses 115
Among the many special and unique features of radiochemistry compared to con-
ventional chemistry is the correlation between an unstable isotope´s radioactivity
level and its minuscule mass. The number of radioactive isotopes Ń is related to the
molar mass of a chemical element: 6.022⋅1023 atoms (nuclides) represent 1 mole, as
per Fig. 5.6. Subsequently, the mass of a given number of radioactive isotopes can
be calculated. 1 mole of 12C involves 6.022⋅1023 nuclides, and the mass of this one
mole is 12.000 g.
Let us apply this correlation to calculate the real mass of a typical level of radio-
activity of 11C and 14C, again for example Á = 1 MBq. The mass of this 1 MBq sample
varies significantly for different isotopes and molar masses: 1 MBq of 11C contains
only 1.7⋅109 nuclides. This is an incredibly small fraction of a mole, namely only
2.9⋅10−15 moles. Mass-wise, it corresponds to the ultra-low amount of 3.2⋅10−14 g
only, which is 3.2⋅10−11 mg or 3.2⋅10−8 μg or 3.2⋅10−5 ng or 320 pg. For the much
longer-lived 14C, a pure 1 MBq sample still only weights 6 μg,7 yet this mass is al-
ready larger by about eight orders of magnitude compared to 11C.
Remember that “normal” preparative chemistry is used to handle chemical compounds at molar
scale, which are typically analyzed gravimetrically. This appears to be very difficult if not impossi-
ble in the case of radioactive substances. In many cases, radiochemists are thus handling radioac-
tive compounds of tiny, almost negligible mass, which cannot be seen by the human eye and are
difficult to analyze by conventional chemical methods!.
116 5 From stable to unstable nuclides: Mathematics
Figure 5.8 shows an initial upper line of 16 orange circles, each representing a
number (e.g. 106 nuclides per circle)8 of unstable nuclides ✶K1. The kinetics could be
illustrated similar to Fig. 5.4, but adapted to the value of t½ = 1.0 h over four hours.
Ń n t½
(t½ = 1 h)
16 0h
8 1h
4 2h
2 3h
1 4h
Fig. 5.8: Example of the statistics of transformation of unstable nuclides. The upper line shows 16
orange circles, each representing a huge number (e.g. 106 nuclides per circle) of unstable nuclides
of one sort ✶K1 with a half-life of t½ = 1.0 h. Within one half-life, e.g. after 1 h, half of these
nuclides have transformed into a stable nuclide K2 (black circle). After 4 h, there is only one orange
circle remaining. It represents the 106 ✶K1 nuclides still in existance. (Actually, the effect is
simplified because nuclear transformations will proceed among all the 106 nuclides of each circle
at the same time).
This corresponds to the common situation in radiochemistry, which is the handling of radioactiv-
ities representing large numbers of nuclides. Figure 5.7 indicates a range of half-lives typically han-
dled in research institutions, i.e. at half-life levels ranging from minutes to years. The typical
numbers of radioactive nuclei (for the Á = 1 MBq level selected) ranged from Ń = 107 to Ń = 1016. For
Fig. 5.7, 16 times the circles of 106 nuclei per circle just made Ń = 1.6⋅107.
5.2 Correlations between radioactivity, number of unstable nuclides and masses 117
9 This is the concept behind the virtual experiment of SCHRÖDINGER’s cat. It was originally meant to
showcase the ideological differences between quantum physics and conventional physics. It is cen-
tered on the fate of a single unstable (radio)nuclide, which is packed together with a cat and a poi-
son vial in a box. The box should remain closed for a certain period of time (which may be
understood as e.g. one half-life of the particular radionuclide). The radionuclide’s transformation
would induce a signal, releasing the poison – and the cat would die. When the box is closed, the
fate of the cat is unknown. However, when the box is opened after a certain time, you can observe
whether the cat is dead or alive. So: there is nothing to predict.
Courtesy: http://commons.wikimedia.org/wiki/File:Catexperiment.svg?uselang = de
Or let us take an ultimate example: two unstable nuclei. When a measurement is performed at
time-point t, if the two nuclei are both still in the form of ✶K1, one would say that they are stable. In
contrast, suppose one nucleus had transformed; its half-life would then equal the time at which the
measurement was taken.
118 5 From stable to unstable nuclides: Mathematics
Fig. 5.9: Different approaches to half-lives. (left) One unstable (orange) nuclide transforms into one
stable (black) nuclide. Since the transformation of unstable nuclides occurs randomly, it is hard to
predict at what time it will occur. (right) The same is true for each unstable nuclide among a larger
sample. However, the signal from many transformation events, which can be measured at any
time-point, provides experimental values of the half-life (or transformation constant). This in turn
describes the probability of these many nuclides to transform over any unit of time.
Conventional chemical reactions and their equilibrium constants deal with a large
number of reagent atoms and/or molecules. The relationship between radioactivity
and mass thus becomes very important and sometimes is not following conven-
tional chemistry. The ultra-low mass and concentration of radioactive compounds,
especially for artificial and short-lived radioelements, require sophisticated experi-
mental approaches.
Important effects are the differences in chemical reaction pathways due to ultra-
low concentration. Halogens, for example, do not exist as a single atom in their ele-
mental form, but as diatomic molecules of type X2. They disproportionate such that
2 HX ⇄ 2 X− + 2 H+ or synproportionate according to X− + HOX + H+ ⇄ X2 + H2O, for
example. While this is true for all naturally occurring fluorine, chlorine, bromine
and iodine, it does not necessarily apply to the heaviest homologue, astatine. This
artificially produced radioelement with the longest-living isotope 211At (t½ = 7.2 h)
exists exclusively at ultra-low concentration. An activity of 1 MBq for 211At corre-
sponds to 3.77⋅10+8 atoms only (<5⋅10−17 mol). At such a low concentration, it is not
able to form a 211At2 molecule because the probability of two single 211At atoms meet-
ing by chance is almost zero; it cannot disproportionate nor synproportionate for the
same reason.
5.3 Units of radioactivity 119
Other features concern reaction kinetics, which are – unlike analogous conven-
tional chemical reactions – typically pseudo first-order in nature.11 These and many
more special aspects are covered by the science of radiochemistry.
The unit of radioactivity is 1/second. The SI unit of radioactivity (since 1975) is the
Becquerel: 1 Bq = 1 s−1. Although the use of SI units is recommended, there are
some cases in natural sciences where “older” units are used in parallel. Since the
discovery of radioactivity by HENRY BECQUEREL in 1896, several other units have
been suggested. The most popular alternative (still commonly used today) is the
“Curie”, (Ci), named after the pioneering contributions of MARIE SKŁODOWSKA CURIE
to radiochemistry and nuclear sciences.12 Historically (1930), it was derived from
the radioactivity represented by exactly 1 gram of the isotope 226Ra of the chemical
element radium (one of the two new elements discovered by M. CURIE). According to
Fig. 5.2 the radioactivity of 1 g of 226Ra (t½ = 1600 a) corresponds to 3.7⋅1010 s−1.
Thus, 1 Ci = 3.7⋅1010 s−1 (or 1 Ci = 37⋅109 Bq = 37 GBq). This unit is more convenient
for quantifying large levels of radioactivity.
The SI unit of Bq expresses activities in rather large numbers. For example, 1 Ci =
37⋅109 Bq, or vice versa: 1 Bq = 2.7027⋅10−10 Ci, 1 MBq = 27.027 μCi. Typical levels of
radioactivity in radiochemical laboratories may thus be given in either multiples of
kBq, MBq or GBq or in multiples of μCi or mCi. This relationship is illustrated in
Fig. 5.10.
The velocity (i.e., the reaction rate R of a chemical reaction) typically correlates the concentra-
tions of e.g. two reactants A and B with a reaction rate constant k. The units of the latter depend on
the reaction order: zero, first, second. For a chemical process of e.g. aA + bB ! cC with given con-
centrations of the species A, B and C and their stoichiometric coefficients a, b and c, respectively,
the rate equation is R = k [A]a [B]b. Over the course of the reaction, the concentration of A decreases,
following −d[A]/dt = k[A]a[B]b. In the case where a = b = 1, it is −d[A]/dt = k[A][B]. Here, both con-
centrations must be considered, making the kinetics second-order. For a (pseudo) first-order reac-
tion, the concentration of radionuclide reactant (A) is extremely low compared to non radioactive
species (B). Due to its excess, the concentration of B does not change remarkably. The new rate
equation is −d[A]/dt = k´[A] with R = k´[A] and k´ reflecting a parameter of k´ = k[B].
Both units thus honor the groundbreaking contributions of the pioneering scientists who were
awarded the Nobel Prize in 1903. The Nobel Prize in Physics, 1903, was divided: one half awarded
to BECQUEREL “in recognition of the extraordinary services he has rendered by his discovery of spon-
taneous radioactivity”, the other half jointly to PIERRE CURIE and MARIE SKŁODOWSKA CURIE “in recog-
nition of the extraordinary services they have rendered by their joint researches on the radiation
phenomena discovered by Professor HENRI BECQUEREL”.
120 5 From stable to unstable nuclides: Mathematics
Fig. 5.10: Units of radioactivity. The SI unit is the Becquerel: 1 Bq = 1 s−1. In parallel, the unit Curie
is used: 1 Ci = 3.7⋅1010 s−1. Both units correlate by 1 Bq = 2.7027⋅10−10 Ci. It is practical to treat it
as 1 mCi = 37 MBq, or vice versa: 1 MBq = 27.027 μCi. The Curie may be better suited to describe
large levels of radioactivity; the Becquerel is more convenient for smaller levels.
The challenge is to correct for all potential losses in counting the initial events.14 In
the case of radiation measurements, several coefficients are considered, addressing
losses of count rate caused by the aforementioned effects. All types of radioactive
Fig. 5.11: Absolute radioactivity vs. experimental count rate. The count rate will be affected by
experimental design, including (among other things) the distance between source and detector,
the detector efficacy (its material, dimension etc.) and the detector shielding. Decreasing intensity
of the black color indicates decreasing intensity of radiation emitted from the source with
increasing distance. Note that the effect shown here in a two-dimensional situation is in reality of
4π geometry. This also applies to the solid angle, which relates to the dimension of the detector
surface. The larger the surface, the larger the solid angle, the larger the fraction of radiation
arriving at the detector.
– the effect of the branching of various types of emissions per one event of nu-
clide transformation (εemission);
– the efficacy with which the detector registers each radiation (εdetector), and;
– the effect of distance between the source and detector, as well as their dimen-
sions (εgeometry).
For the latter coefficient, for example, the closer the source and detector are, the
higher the εgeometry. Each coefficient ranges from 0 to 1, but is typically far below 1.
Tab. 5.4: Absolute activity vs. count rate, and absolute activity vs. specific activity.
Parameter Units
−λt
Count rate Ć = Ćo e cps, cpm
dependent on the experimental design
(detector geometry, distance, . . . )
Á
Ás = P (5:12)
mi
Consider the following example. The activity of 1 MBq of artificially produced 11C
corresponds to 3.2⋅10−14 g, as shown in Tab. 5.3. Suppose this fraction is in contact
with air, plastic tubes, glass vessels, chemical compounds etc., which usually are
composed of carbon to a significant part, then 11C and 12C+13C will mix. Figure 5.12
takes a (low16) mass of 1 μg of natural carbon, which co-exists with 11C. What about
the specific activity now? The overall mass of carbon atoms would increase signifi-
cantly from 3.2⋅10−14 g to 3.2⋅10−14 + 1⋅10−6 g = 1.000000032⋅10−6 g. The specific
activity would decrease by seven orders of magnitude.
Á(11C)= 1 MBq
Ás =
Ń + Ń(stable1) + Ń(stable2)
Fig. 5.12: “Real” specific activity of an 11C sample containing altogether 1 μg of stable carbon. The
corresponding masses of 12C and 13C (with the 1.0 μg attributed to the two stable carbon isotopes
according to their percentages of 98.93:0.107) yield a real specific activity of 1⋅10+6 Bq /
1.000000032⋅10−6 g ≈ 1⋅10+12 Bq/g. The theoretical specific activity of 11C according to eq. (5.5) is
THEORY
Ás = 3.15⋅1019 Bq/g (or 8.4⋅108 Ci/g).
Table 5.5 summarizes special categories used in terms of specific activity related to
mass. The categories are “carrier free”, “no carrier added”, and “carrier added”. The
wording “carrier” refers to any amount of stable isotope(s) of a chemical element of a
specific mass, which is much higher than the mass of the unstable isotope (Ń ≪
Ń(STABLE1) + Ń(STABLE2) + . . . .), in particular for relatively short-lived unstable nuclides.
Until now the transformation of an unstable nuclide (✶K1) into a stable nuclide (K2)
was discussed within one transformation step: ✶K1UNSTABLE ! K2STABLE. However,
the process of stabilizing an unstable nucleon configuration may proceed in several
steps, and the first transformation may result in a “more stable”, but not “fully sta-
ble”, nuclide. Thus, this first transformation yields a second unstable nuclide ✶K2,
which undertakes a further transformation step to obtain stable nucleon composi-
tion: ✶K1UNSTABLE ! ✶K2 UNSTABLE ! K3STABLE. This constellation is particularly rele-
vant since the first radioactive nuclide “generates” a second radioactive one. This
intermediate is both being formed and transformed simultaneously. The constella-
tion is called “equilibrium” and in cases where the intermediate nuclide ✶K2 is of
practical interest, it is referred to as “radionuclide generator”.
Most of the unstable nuclides, however, are so far from having stable nucleon
compositions that several transformation steps are needed until a final, stable con-
figuration is reached. This is called a “transformation chain” (or “decay chain”).
Some of the intermediate nuclides may undergo different transformations in paral-
lel. In this case, an initial chain branches into two or more parallel chains.
The three types mentioned are illustrated in Fig. 5.13.
5.4 Classes of radioactive transformations 125
SINGLE TRANSFORMATION
K1 1 K2
EQUILIBRIUM
K1 1 K2 2 K3
CHAIN
K1 1 K2 2 K3 3 K4 4 K5 6 K6 8
9
7
5
BRANCHED CHAIN
K7 10 K8 11 K9 12
Fig. 5.13: Classes of radioactive transformation, which require separate mathematical treatment.
White and black boxes denote unstable and stable nuclides, respectively. The simplest case is to
transform one unstable nuclide into a stable one: ✶K1 ! K2. There is one transformation constant
only: λ1. If the initial process yields one unstable intermediate ✶K2, a second transformation step is
required to obtain stable nucleon composition: ✶K1 ! ✶K2 ! K3. This involves two different
transformation constants, λ1 and λ2. For a “chain”, there may be a number of successive
transformation steps with their individual transformation constants. Some of the intermediate
nuclides may undergo different transformations in parallel, for example at stage ✶K4. Here, two
different transformation types proceed simultaneously, and two different unstable nuclides ✶K5
and ✶K7 are formed with specific λ4 and λ10 constants, respectively. Both continue to transform. In
this case, an initial chain which started from ✶K1 continues along two (or more) parallel “branched
chains”.
d✶ Ń1 dŃ2
− =+ (5:13)
dt dt
126 5 From stable to unstable nuclides: Mathematics
100
90
80
K2
70
60
Ń (%)
50
40
30
20
*K1
10
0
0 1 2 3 4
t/t1/2
Fig. 5.14: Kinetics of radioactive transformation. (Ń = Ńo e−λ₁t). An unstable nuclide ✶K1 “turns” into
a stable nuclide K2. After one half-life of the nuclide ✶K1, 50% of unstable ✶K1 had transformed into
stable K2.
5.4.3.1 Equilibrium
The wording “equilibrium” refers to the situation that the two radioactivities of ✶K1
and ✶K2 reach a certain ratio that remains constant over time: the radioactivity of
✶
K2 is in “equilibrium” with the radioactivity of ✶K1. “Equilibrium” will occur if the
half-life of ✶K1 is larger than that of ✶K2. More precisely, two subtypes may be dis-
cussed: a secular (λ 1 ≪ λ2) and a transient (λ1 < λ2) equilibrium. (The same is true
when considering half-lives, and the cases will be t½(1) ≫ t½(2) or t½(1) > t½(2).)
Mathematically, there are relations of type λ1 = x λ2 with x > 1. In order to decide be-
tween the two categories λ 1 ≪ λ 2 or λ1 < λ2, one may fix the ratio according to the
factor x as either x > 100 or 1 < x < 100, as shown in Table 5.6.17
no TRANSIENT SECULAR
equilibrium equilibrium equilibrium
Fig. 5.15: Categories of transformation of type ✶K1 (via λ1) ! ✶K2 (via λ2) ! K3, related to the ratio
between transformation constants. The two equilibria are categorized according to λ1 being smaller
than λ2 for ratios of λ1:λ2 ranging from 1:1 to 1:100 (transient) and λ1:λ2 being larger than 1:100
(secular).
Tab. 5.6: Versions of radioactive equilibrium of type ✶K1 ! ✶K2 (via λ1)
and ✶K2 ! K3 (via λ2).
The case of λ1 < λ2 would mean that the value of t½(1) is “a bit” greater than the one for t½(2).
This appears to be somehow arbitrary, as illustrated in Fig. 5.14, because there is actually no defini-
tive borderline between the ratios of the two constants.
128 5 From stable to unstable nuclides: Mathematics
*Ń1 = *Ń1(o) e- t
1 2
d*Ń2
= * *
+ 1 Ń1 – 2 Ń2
dt
4
d*Ń2
= * * e- t
2 Ń2 – 1 Ń1(o) =0
dt
Fig. 5.16: Mathematical expressions for transformation equilibria of type ✶K1 (via λ1) ! ✶K2 (via λ2)
! K3. The first step reflects the known exponential law for the transformation of an unstable
nuclide ✶K1. Step 2 thus says −d✶Ń1/dt = + d✶Ń2/dt. Each nuclide of ✶K2, once formed, transforms
into the stable nuclide K3. Step 3 describes the kinetics of ✶K2 as simultaneously being formed
and decomposed. In phase 4, the number of Ń2 (which changes over time) is in “equilibrium” with
the two processes of formation and transformation. This depends on the individual transformation
constants λ1 and λ2 and more specifically, on the ratio between λ1 and λ2, as discussed below.
Each nuclide of ✶K2, once formed, transforms into a stable nuclide of K3 via λ2. The
overall kinetics then follow eq. (5.14) in a differential form of d✶Ń2/dt = + λ1✶Ń1 − λ2✶Ń2,
i.e. as formation and transformation (decay). Equation (5.15) expresses the same con-
cept, now with a specific number of unstable nuclides ✶K1 known at a given time-
point t = 0. With the initial number of ✶Ń1(o) known at any time-point t = 0, the change
in ✶Ń2 over time (which is radioactivity) is obtained.
d✶ Ń2
= + λ 1 ✶ Ń1 − λ 2 ✶ Ń2 (5:14)
dt
d✶ Ń2
= + λ 1 ✶ Ń1ð0Þ e − λ 1 t − λ 2 ✶ Ń2 (5:15)
dt
+ λ 1 ✶ Ń1 = − λ 2 ✶ Ń2 (5:16)
d✶ Ń2
0= + λ 2 ✶ Ń2 − λ 1 ✶ Ń1ð0Þ e − λ 1 t (5:17)
dt
Since eq. (5.17) is a differential one, integration results in eq. (5.18). The main mes-
sage is: with the initial number of ✶Ń1 known at time-point t = 0, the number of ✶Ń2
is mathematically derived according to the time-point t considered and the values
of the two transformation constants λ1 and λ2. While eq. (5.18) refers to a situation
in which there were no nuclides of ✶K2 present at t = 0, eq. (5.19) considers this
case, that a certain number ✶Ń2 of nuclides of ✶K2 was present, as per Fig. 5.17. The
total number of ✶Ń2 existing is thus composed of those that existed earlier (i.e. at
t = 0), and the new ones generated by the transformation of ✶K1.
me-point
transforming *Ń2
nts
Ń2
Ń1
generating *Ń
decay constan
into stable Ń3
present at tim
number of *Ń
considered
out of *Ń1
ratio of
1
*Ń2 = *Ń1(o) (e- 1t – e- 2t)
2- 1
1
*Ń2 = *Ń1(o) (e- 1t – e- 2t) + *Ń2(o) e- 2t
2- 1
Fig. 5.17: Integral version of eq. (5.17). With the initial number of ✶Ń1 known at time-point t = 0, and
with the number of ✶Ń2 at this moment being zero, the number of ✶Ń2 is mathematically derived
according to the time-point t considered and the values of the two transformation constants λ1 and
λ2. There are four terms which influence the number and the kinetics of ✶Ń2. First, of course, it is
the number of ✶Ń1 existing at time-point t = 0. Next, is the ratio of the two transformation
constants, namely λ1/(λ2 – λ1) and their individual contributions according to (e− λ 1 t − e− λ 2 t ). Finally,
eq. (5.19) considers the case where the total number of ✶Ń2 existing are composed of new ones
generated by the transformation of ✶Ń1 plus the ones that existed earlier (i.e. at t = 0) and which
transform separately via ✶Ń2(o) e− λ 2 t .
130 5 From stable to unstable nuclides: Mathematics
✶ λ1 ✶
Ń2 = Ń1ð0Þ e − λ 1 t − e − λ 2 t (5:18)
λ2 −λ1
✶ λ1 ✶
Ń2 = Ń1ð0Þ e − λ 1 t − e − λ 2 t + ✶ Ń2ð0Þ e − λ 2 t (5:19)
λ2 −λ1
There are several tasks related to the mathematical description of this type of trans-
formation and in particular to the activity of ✶K2 available at various moments. Sup-
pose the initial values for Á1 and ✶Ń1 are known:
1. At any time: What is the radioactivity of ✶K2 given the initial activity of ✶K1?
2. At equilibrium: How does the radioactivity ✶K2 relate to the radioactivity of
✶
K1?
3. At what time is the maximum activity of ✶K2 available in the system?
The exact calculation is based on eq. (5.18). However, in the case of either λ1 < λ2 or
λ1 ≪ λ2, two simplifications will modify eq. (5.18). The concept is illustrated in
Fig. 5.18.
Á1 2-
(5.20)
1 *Ń1 2- 1 1
*Ń2 = *Ń1 = =
2- 1 Á2
*Ń2 1
2
e- t « e- t
1 < 2: TRANSIENT
*Ń1 = *Ń1(o) e- t
1
Ń1(o) (e- 1t – e- 2t)
8)
**Ń
Ń2 = **Ń
(5.18
2- 1
2 - 1 2 1 « 2: SECULAR
*Ń1
(5.21)
1 *Ń1 2
*Ń2 = = Á1 = Á2
*Ń2 1
2
Fig. 5.18: Two simplifications of the general equation of transformation equilibria of type ✶K1
(via λ1) ! ✶K2 (via λ2) ! K3. When λ1 < λ2, it yields eq. (5.20) due to the simplifications of
(e − λ 2 t ≪ e − λ 1 t ) ≈ e − λ 1 t and ✶Ń1(o) e − λ 1 t = ✶Ń1. When λ1 ≪ λ2, this continues to simplify due to
λ1/(λ2 − λ1) ≈ λ1/λ2 and yields eq. (5.21).
5.4 Classes of radioactive transformations 131
✶ λ1
Ń2 = ✶ Ń1 (5:20a)
λ2 − λ1
✶ λ2
Á2 = ✶ Á1 (5:20b)
λ2 −λ1
1 λ2
t✶ K2max = ln (5:20c)
λ2 −λ1 λ1
Figure 5.19 illustrates the most relevant features of a transient transformation equi-
librium for a situation in which an initial number ✶Ń1(o) of unstable nuclides ✶K1 is
present without any nuclides of ✶K2. Let the ratio between the transformation con-
stants be 10, i.e. λ2 = 10ˑλ1, indicating the half-life of ✶K1 is longer than ✶K2 by a fac-
tor of 10. The difference in the exponential form becomes significant: e − λ 2 t ≪ e − λ 1 t
and e − λ 2 t may be neglected. ✶Ń1(o)e − λ 1 t remains and is written as ✶Ń1.
From t = 0 on, ✶K1 starts to transform according to λ1 and generates ✶K2. Within
the 10 periods of the half-life of ✶K2, which is exactly 1 half-life of ✶K1, the value of
Á1 drops to 50%.
Á2 is defined by both continuous formation ✶K1 ! ✶K2 and subsequent transfor-
mation ✶K2 ! K3. Each ✶K2 generated adds a given activity Á2 on top of Á1. Both
values of activity combine to a cumulative Á1 + Á2, which is larger than either Á1 or
Á2 alone. After several periods of half-life, the cumulative activity parallels Á1 – fi-
nally meaning what? ✶K2 is “in equilibrium” with ✶K1.
Equation (5.20c) allows for the time-points of two maxima to be determined;
one for the cumulative activity, and the other (more importantly) for the maximum
activity Á2max. This parameter can be very important in the practical use of ✶K2,
since it may define a time-point when ✶K2 is radiochemically separated from the
mixture of ✶K1 + ✶K2. Once isolated, it will transform individually according to its
own half-life.
Ámax(*K1+*K2)
3
100
1
Ámax
2 (*K2)
10
4 5 6
1
0 1 2 3 4 5 6 7 8 9 10
t/t½ (*K2)
Fig. 5.19: Kinetics of transient equilibrium showing activities of ✶K1 and ✶K2 depending on the
period of half-lives of ✶K2. The ratio between the transformation constants is fixed to x = 10, i.e.
λ2 = 10ˑλ1. (1) Transformation of ✶K1 according to λ1 expressed at Á1, (2) increasing activity Á2
starting at Á2 = 0 at t = 0. This corresponds to eq. (5.12) and would be different if a given fraction of
✶
K2 ws already present at t = 0, as expressed in eq. (5.19) by + ✶Ń2(o)e − λ 2 t . (3) Both values of
activity combine as Á1 + Á2, and this cumulative activity is larger than either Á1 or Á2 alone. (4) The
time-point of maximum value for cumulative activity; in this case, 2.637 t/t½. (5) Maximum activity
of Á2 achievable in the equilibrium according to t*K2max = [1 / (λ2 – λ1)] ln(λ2/λ1), eq. (5.20c);
3.697 t/t½ in this case. (6) Once ✶K2 is radiochemically separated from the mixture of ✶K1 + ✶K2, it
will transform according to its own half-life.
term λ1 / (λ2 − λ1) is approaching λ1/λ2 because λ1 is very small compared to λ2.18 For
the two nuclides ✶K1 and ✶K2 at equilibrium, the relationships simplifies according
to eqs. (5.21a) and (5.21b):
For example, a ratio of λ1:λ2 of 1:1000 would be either 1 / (1000 − 1) in the original version of
eq. (5.20) or 1 / 1000 in the version of eq. (5.21). The two values of 0.001001001 and 0.001000000
differ by about 0.1% only. This is not the case for e.g. a ratio of λ1:λ2 of 1:10. This is the reason to
classify transient and secular systems by ratios of λ1:λ2 of 1 to 100 and >100, respectively.
5.4 Classes of radioactive transformations 133
✶ λ1
Ń2 = ✶ Ń1 (5:21a)
λ2
Á2 = Á1 (5:21b)
Figure 5.20 illustrates the kinetics of secular transformation equilibrium. Let the
ratio between the transformation constants be x = 1000, i.e. λ2 = 1000λ1, indicating
the half-life of ✶K1 is longer than ✶K2 by a factor of 1000. Again, an initial number
✶
Ń1(o) of unstable nuclides ✶K1 is present without any nuclides of ✶K2, and starts to
transform according to λ1 generating ✶K2. This time, the activity Á1 does not change
significantly over the time period considered. After ten periods of t½(✶K2), just
0.1% of one half-life of t½(✶K1) have passed, and the fraction of ✶Ń1 still present
(according to 1/2n) is 99.309%.
Each ✶K2 generated adds a given activity Á2 on top of Á1. After several half-life
periods, the cumulative activity again parallels Á1. At equilibrium both activities
are the same: Á2 = Á1. Both values of activity combine to a cumulative Á1 + Á2,
which is exactly double the individual values of Á1 or Á2. Again, ✶K2 is “in equilib-
rium” with ✶K1, but this time the individual activity Á2 also remains (practically)
constant – at least for the 10 half-lives of ✶K2 shown in Fig. 5.20. Unlike the transient
system, no true maxima are reached, neither for the cumulative activity, nor for the
maximum activity Á2 (the latter is a saturation effect). Starting from t = 0, the fraction
of ✶K2 generated will exponentially increase. After one half-life, this is 50%, after two
half-lives it is 75%, after three half-lives 87.5%. The activity Á2 generated approaches
a maximum (asymptotic) level, and then will remain at this value.
For practical applications, ✶K2 may radiochemically be separated from the mix-
ture of ✶K1 + ✶K2 at a time-point when a sufficient value of Á2 is generated. This
could be a reasonable time-point to separate ✶K2. Once isolated, it will transform
individually according to its own half-life.
5.4.3.6 No equilibrium
Let’s assume the ratio between the transformation constants is λ2 = 0.1ˑλ1, i.e. the
half-life of ✶K1 is shorter than ✶K2 by a factor of 10. Figure 5.21 illustrates the trans-
formation kinetics in this case. Again, ✶K1 starts to transform according to λ1 and
generates ✶K2. Both values of activity combine to a cumulative Á1 + Á2, which is
larger than either Á1 or Á2 alone. However, the cumulative activity never parallels
Á1 – this means there is “no equilibrium”.
3 SECULAR
EQULIBRIUM
100 1
10
1
0 1 2 3 4 5 6 7 8 9 10
t/t½ (*K2)
Fig. 5.20: Kinetics of secular equilibrium showing activities of ✶K1 and ✶K2 depending on the
period of half-lives of ✶K2. The ratio between the transformation constants is fixed to x = 1000, i.e.
λ2 = 1000ˑλ1. (1) Transformation of ✶K1 according to λ1 expressed at Á1, staying practically constant
for the 10 periods of half-life of ✶K2 considered. (2) Activity Á2 generated starting at Á2 = 0 at t = 0.
This would be different supposing that at t = 0 a given fraction of ✶K2 would already have been
present in the system, as expressed in eq. (5.19) by + ✶Ń2(o)e − λ 2 t . At equilibrium, Á2 approaches
Á1. (3) Cumulative activities at equilibrium are Á = Á1 + Á2, and this cumulative activity is larger by a
factor 2 than either Á1 or Á2. (6) Once ✶K2 is radiochemically separated from the mixture of ✶K1 + ✶K2,
it will transform individually according to its own half-life. Starting from t = 0, the fraction of ✶K2
generated relative to its maximum value will exponentially increase.
generator”.19 The longer-lived “precursor” may be installed at any laboratory and pro-
vide the shorter-lived offspring “on demand”. Thus, it is radiochemical expertise that
However, the number of generator systems available is rather limited. This is mainly due to the
fact that nuclear transitions of type ✶K1 ― λ1 ! ✶K2 ― λ2 ! K3 with λ1 < λ2 are rather “atypical”. As
differences in nuclear mass, mass excess and mean nuclear binding energy become smaller for
transitions towards the finally stable nucleus, the half-lives of the successive nuclei increase and
the transformation constant decreases. Thus, the “typical” situation for ✶K1 ― λ1 ! ✶K2 ― λ2 ! K3 is
λ1 < λ2 and t½(✶K1) > t½(✶K2). See the detailed discussions of β- and α-transformation processes.
5.4 Classes of radioactive transformations 135
100
1
10
Á
2 6
0,1
0 1 2 3 4 5 6 7 8 9 10
t/t½ (*K2)
Fig. 5.21: No equilibrium is reached for transformation of type ✶K1 (via λ1) ! ✶K2 (via λ2) ! K3. Let the
ratio between the transformation constants be x = 0.1, i.e. λ2 = 0.1ˑλ1, iindicating the half-life of ✶K1 is
shorter than ✶K2 by a factor of 10. (1) Activity Á1 changes significantly over the time period considered.
After ten periods of t½(✶K2), 100 half-life periods of ✶K1 have passed. The fraction of ✶Ń1 still present
is 1/2n = 1/2100 = 7.9⋅10−31. This is negligible. (2) Á2 is defined by both continuous formation ✶K1 !
✶
K2 and subsequent transformation ✶K2 ! K3. (3) Cumulative activity Á1 + Á2 has no equilibrium.
(6) Activity Á1, once ✶K2 has been radiochemically isolated from the mixture of ✶K1 + ✶K2.
turns a formal transformation system of type ✶K1 (via λ1) ! ✶K2 (via λ2) ! K3 into a
“generator” of practical relevance. The two different unstable nuclides ✶K1 and ✶K2
most often represent isotopes of different chemical elements. When designing a system
to exploit the significant differences in chemical behavior between the parent and
daughter nuclides (elements), the goals are threefold:
1. Achieve high separation efficacy when isolating the chemical fraction of ✶K2
from the chemical fraction of ✶K1. This is called “separation” yield.
2. Achieve low turnover of parts of ✶K1 still present in the separated fraction of
✶
K2. This is called “breakthrough”.
3. Guarantee long shelf life of the generator system, allowing many subsequent
separation cycles of ✶K2.
136 5 From stable to unstable nuclides: Mathematics
Table 5.7 lists some of the radionuclide generator systems of practical relevance.
Tab. 5.7: Some of the radionuclide generator systems of practical relevance, arranged according to
increasing ratios of t½(✶K1) / t½(✶K2).
The last means of transforming unstable, radioactive nuclides into more stable ones
considers nuclides with mean nucleon binding energies far from being stable. Here,
transformation proceeds stepwise, including formation of many intermediate unsta-
ble nuclides. The most prominent examples of transformation chains are the so-
called “naturally occurring decay chains”. Theoretically, there are four chains, which
originate from four individual and long-lived unstable heavy nuclides. Table 5.8 sum-
marizes the four “parent” nuclides, all belonging to the actinide elements: one tho-
rium isotope, one neptunium isotope, and two uranium isotopes: 232Th, 237Np, 238U
and 235U. They are located near the very end of the Chart of Nuclides and form an
“island” of semi-stable nuclides, as per Fig. 5.22.
The final destinations of the four chains are four corresponding individual stable
nuclides, namely 208Pb, 209Bi, 206Pb and 207Pb, respectively. These final nuclides are
See Chapter 7 on Radiochemical separations and Chapter 13 on Molecular imaging, Vol. II.
5.4 Classes of radioactive transformations 137
characterized by their high mean nucleon binding energy, which are additionally
stabilized due to magic nucleon numbers: Z = 82 for the lead isotopes, N = 126 for
the bismuth isotope. Every chain covers many unstable nuclides located between
A = 232–238 and A = 206 − 208.
Sometimes each chain is called a “family”. (Remember the case of the two-step
transformations creating radionuclide generators: the first unstable nuclide ✶K1
again is a “parent”, the subsequent unstable nuclide ✶K2 is a “daughter”.) Similarly,
232
Th, 237Np, 238U, and 235U induce daughters and granddaughters etc., thereby
forming a genetically consistent family. The “genes” in this case lie in the mass
number A, and the four chains are unique in forming follow-up nuclides character-
ized by mass number of A = 4n + i, with i being 0, 1, 2, or 3 and thus individual for
each “family”. For example, “daughters” of 238U belong to the (i = 2) family. 238U it-
self is A = 4 × 59 + 2 = 238. Corresponding daughter nuclides must have mass num-
bers of A = 4 × 58 + 2 = 234. Next-generation transformation creates nuclides of mass
94
237Np
93
2.14.106 a
235U 238U
92
7.04.108 a 4.47.109 a
91
Z
232Th
90
1.41.1010 a
89
88
87
Fig. 5.22: Position of the four parent nuclides of the naturally occurring transformation chains in
the Chart of Nuclides. Each of the four nuclides starts with an α-transformation and continues with
α- and/or β-transformations.
138 5 From stable to unstable nuclides: Mathematics
number A = 4 × 57 + 2 = 230 and so on. The final, stable nuclide is of mass number
A = 206, which is 4 × 51 + 2, representing 206Pb.21
What does a “naturally occurring” chain mean? It may mean that the transforma-
tions are “natural”, going on since the planet earth was formed, ca. 4.55⋅109 years
ago. Well, this is true – more or less. Table 5.9 compares the half-lives of the four
parent nuclides with the age of the earth. The fraction of unstable parent nuclides
still present after a number n of half-lives is 1/2n. If related to the age of the earth, a
fraction of n´ may be defined as the ratio of n´ = tearth/t½(✶K1) and the fraction still in
existance is 1/2 n´. In the case of 238U, for example, its half-life is almost the age of the
earth and n´ = tEarth/t½ = 4.55⋅109 a / 4.47⋅109 a = 1.0179. Since 1/21.0179 is 0.492835,
about 49% of 238U nuclides are still present on earth today compared to earth´s gene-
sis ( = 100%). For 232Th, it is even more: 1/20⋅3,238,434 = 0.7989386, i.e. 79.9%. These
two nuclides form daughters and granddaughters – naturally, during the past 4.55⋅109
years and still today. The situation is qualitatively the same for 235U, but quantitatively
different: with n´ = 6.4649 for 235U, only 1.132% is still present. However, with n´ > 2000
for 237Np, this nuclide has expired – it no longer naturally exists on earth anymore.
Consequently, there is no longer a naturally occurring 237Np transformation chain.
All four chains follow the characteristics introduced in Fig. 5.13. For each chain,
there are some parts of type ✶K1 (via λ1) ! ✶K2 (via λ2) ! ✶K3 (via λ3) ! ✶K4 (via λ4)
! and so on, but there are also regions where the chain branches.22 Let us take the
The understanding of chain transformations in the context of a “family” led historically to spe-
cific notations for some family members. At times, the individual daughter nuclides have been identi-
fied in terms of emission parameters such as half-life rather than chemical parameters, successive
transformation products were named according to the parent. For example, some of the transforma-
tion products of 226Ra were considered “post-radium-nuclides” by e.g. RaA ( = 218Po), RaB ( = 214Pb),
RaC ( = 214Bi), and, because 214Bi branches, even RaC´( = 214Po), and RaC´´ ( = 210Tl), etc.
The mathematics involved in generating several unstable nuclides in a row, particularly when
branching chains are considered, are more complex than the equilibrium between two unstable nu-
clides. This mathematics is not discussed here. Instead, the typical performance of a natural trans-
formation chain is described.
5.4 Classes of radioactive transformations 139
238
U chain, for example. Its complete series is illustrated in Fig. 5.23, containing the
excerpt of the Chart of Nuclides ranging from Z = 80 (U) to 93 (Np) and N = 124 to 146.
There are two directions of transformations within the chain. Namely, there is a
long step from upper right to lower left crossing one field within the Chart of Nu-
clides, and a short one from lower right to upper left. These transformations belong
to the emission types of either β− or α; see Chapters 8 and 9.
The linear part of the chain, the isodiaphere line, ranging from 238U to 218Po
and the beginning of the first branching of the chain at 218Po, which simultaneously
transforms into 214Pb and 218At, is highlighted. Four of the following nuclides
branch too, namely 218At, 214Bi, 210Pb, and 210Bi. 226Ra is also highlighted.23
Finally, Fig. 5.24 illustrates the pathways of all the four chains. Among the mem-
bers of the four families, there is no overlap because of the different “genetic code”
of A = (4n + i). Because of the many steps involved when transforming the unstable,
long-lived parent nuclide into the final stable one, the four families populate a sig-
nificant number of intermediate nuclides, showing a broad range of half-lives.
Table 5.10 summarizes all the nuclides involved in the four chains.
Although different from the radionuclide generators of type ✶K1 (via λ1) ! ✶K2
(via λ2) ! K3 described above, radionuclide transformation chains provide family
members of scientific and practical interest that may be isolated radiochemically.
Many of these nuclides became relevant in fundamental nuclear science, signifi-
cantly contributing to the progress in nuclear and radiochemistry, but also in
modern applications in medicine and technology. Some are listed in Tab. 5.11.
One example was already mentioned: the 226Ra isolated by M. CURIE from the 238U
family, which existed in uranium ores at equilibrium. Another important example,
Because this member of the 238U natural transformation chain was the first to be separated from
238
U (according to the equilibrium within the whole chain) and identified as a new chemical ele-
ment. It later became the traditional unit of radioactivity (see above: 1 g of 226Ra = 1 Ci).
140 5 From stable to unstable nuclides: Mathematics
Z= 238U
93
92
U U
91
U
90
U U
89
88
U
87
86
U U 2
85
U
84
U U 1
U
83 U U
82 U U U 3
81 U U
80 U
4
206Pb
Fig. 5.23: The naturally occurring 238U transformation chain. The excerpt from the Chart of Nuclides
ranging from Z = 80 (U) to 93 (Np) and N = 124 to 146 identifies fields for the members of the A =
4n + 2 family. (They all are indicated by “U” to clarify they all are uranium´s daughter’s and
granddaughters.) The final black square is for the stable terminal nuclide, 206Pb. Several aspects
are highlighted: (1) the first part of the chain ranging from 238U to 218Po including α linear line
starting from 234U. (3) The beginning of the first branching of the chain at 218Po, which
simultaneously transforms into 214Pb continuing the α-transformation, and 218At, according to β–
transformation. (4) Four of the following nuclides branch too, namely 218At, 214Bi, 210Pb and 210Bi.
The final stable nuclide 206Pb is thus being populated by 210Po and by 206Tl. (2) 226Ra.
originated from the same chain, was the use of the . . . ! 210
Pb ― t½ = 22⋅3 a
!
210
Bi ― t½ = 5⋅013 d ! . . .24
5.5 Outlook
It was isolated and applied by G. HEVESEY for fundamental physico-chemical and pioneering
medical applications (Nobel Prize in Chemistry, 1943, “ . . . for his work on the use of isotopes as
tracers in the study of chemical processes”).
5.5 Outlook 141
Z=
93
Np
92
Np U Ac U
91
Ac Np U
90
Ac Th Np U Ac Th U
89
Np Ac Th
88
Ac Th Np U Th
87
Np Ac
86
U Ac Th U
85
Ac Th Np U Ac
84
U Ac Th Np U Ac Th U
83
Np U Ac Th Np U Ac
82
U Ac Th Np U Ac Th U
81
U Ac Th Np U
80
U
140
146
144
128
130
138
126
136
124
142
129
134
139
143
132
133
127
145
137
125
135
141
131
N=
Fig. 5.24: Individual pathways of the four chains, starting from the “parents” Ac, U, Th, Np. The
“family tree” of each parent is shown superimposed on the chart of nuclides. There is no overlap
among the members of the families because of the different genetic code of A = 4n + i.
statistics. Key parameters are the half-life t½ or its inverse pendant, the transforma-
tion constant λ.
An important message in the context of radio- and nuclear chemistry is the cor-
relation between the number Ń of radioactive nuclei and their radioactivity Á. It re-
veals that in many cases, a relatively high level of radioactivity is expressed by a
rather low number of nuclei. Mathematics is relevant for various practical applica-
tions including the design and the handling of radionuclide generator systems.
However, the way a nuclide “organizes” its transformation is itself very com-
plex. The different classes of transformation of one unstable nuclide into a more
stable one and the kind of emissions that accompany the nuclear processes are in-
troduced in the next chapters.
142 5 From stable to unstable nuclides: Mathematics
Tab. 5.10: Half-lives of the unstable “daughter” nuclides involved in the four transformation chain
“families”.
Th (A = n) Np (A = n + )
β− β−
Ra . a Pa . days
β−
Ac . h U .⋅ a α
α
Th . a Th .⋅ a α
α β−
Ra . d Ra . days
α α
Rn . s Ac . days
α α
Po . s Fr . min
β−
Pb . h At .⋅− s α
α, β− α, β−
Bi . min Bi . min
Po .⋅− s α
Po . × − s α
β− β−
Tl . min Tl . min
β−
Pb Stable Pb . h
Bi Stable
U (A = n + ) U (A = n + )
β− β−
Th . days Th . h
m β−
Pa . min Pa .⋅ a α), β−
β− α
Pa . h Ac . a
U .⋅ a α
Th . days α), β−
α α
Th .⋅ a Fr . min
α α, β
Ra . a Ra . days
α α
Rn . days At . min
α, β β−
Po . min Rn . s
β− α, β
Pb . min Bi . min
α, β
At ≫s Po .⋅− s β−
− α α
Rn ⋅ s Pb . min
α, β−
Bi . min At .⋅− s α, β
Po . × − s α
Bi . min α
β− α
Tl . min mPo . s
α, β− β−
Pb . a Po . s
β− β−
Hg . min Tl . min
α, β− β−
Bi . days Pb Stable
β−
Tl . min
α
Po . days
Pb Stable
5.5 Outlook 143
Tab. 5.11: Examples of relevant members of the natural transformation chains, radiochemically
isolated from the chain equilibrium.
Aim: Unstable nuclei convert into stable ones by minimizing the absolute mass of a
nuclide, which is typically expressed in terms of nucleon binding energy. The com-
mon feature of all primary transformations is their exothermic character, and the
corresponding energy is introduced in this chapter in a general approach.
There are three principal approaches as derived from the LDM: (a) “simply”
changing the ratio between protons and neutrons, thereby retaining the same mass
number; (b) reducing the mass number by emitting single nucleons or clusters of
nucleons; or (c) dividing the whole nucleus into (mainly) two pieces. The corre-
sponding pathways of transformation are for type (a) the class of β-processes, for
type (b) the α-emission processes and for type (c) spontaneous fission.
The dominant route of transformation, occurring in most unstable nuclides, is
the β-process. Its essence is: turning a neutron into a proton, or vice versa. This class
of transformation involves emission or capture of an electron. Therefore, the mass
number remains the same, and the nuclides involved are isobars. It includes three
individual subtypes: the β−-process, β+-process, and electron capture (EC or ε).
Among the heavy nuclides, α-emission and spontaneous fission become addi-
tional options.
Primary transformations create a new nucleon composition of the nucleus formed,
which does not necessarily represent the ground state of that nucleus. Instead, excited
nuclear levels are populated and these undergo de-excitation through secondary
transitions.
Unstable nuclei convert into stable ones by minimizing their absolute mass, which
is typically expressed in terms of nucleon binding energy. The absolute mass of the
transformation product(s) should be less than the absolute mass of the initial unsta-
ble nuclide. The difference in mass (Δm) is expressed as an absolute mass value,
but typically it is converted into values of energy (ΔE)1 according to ΔE = Δmc2.
Transformations proceed exothermically and, similar to conventional chemical re-
actions, this yields energy: there is less energy “located” in the newly formed nu-
clide compared to the initial, less stable one.
The amounts of energy of these transformation reactions are much higher than energy balances
of conventional chemical reactions.
https://doi.org/10.1515/9783110742725-006
146 6 Processes of transformations: Overview
mnuclide
K1 > mnuclide
K2 (6:1)
Parallel to the changes within the nuclei, one or two small components are ejected
during the transformation. The corresponding subtypes of primary transformation
processes thus also differ in terms of the radiation2 emitted within the course of the
transformation. The primary step of transformation of an unstable nuclide into a
stable (or at least a more stable) one is accompanied by the emission of another
component. It is denoted here as “x” and/or “y”.
The following is valid in all cases:
– spontaneous transformations are exothermic,
– the sum of the mass of the transformation products is less than the mass of the
initial nuclide,
– a characteristic amount of energy is obtained as ΔE.
Figure 6.1 illustrates the general concept. Equation (6.1) summarizes the three as-
pects: change in nucleon configuration from ✶K1 to K2, release of components x
(and/or y), and gain in energy ΔE.
MASS OR ENERGY
*K1 K2 + x + E
Fig. 6.1: Simplified scheme of primary transformation of an unstable nuclide ✶K1 into a stable
nuclide K2.
Mass numbers may remain the same (A2 = A1, the β-processes) or become lower
(A2 < A1, in the case of α-emission and spontaneous fission). Transformations are accompanied by
the emission of one or two components x and y. This “x” is typically a particle, such as a 4He
nucleus (the α-particle) or β-particle (β-processes).
However, eq. (6.2a) is the general representation of the whole process. For the indi-
vidual primary transformation of an unstable nuclide ✶K1 into a stable (or at least a
more stable) nuclide K2, eqs. (6.2b–g) apply. Accordingly, transformation processes
Actually, the transformation itself is the source of what is called “radioactivity” and the unstable
nuclei are the “radioactive” ones. Usually, this type of transformation is called “decay”. The way
“radioactivity” is detected usually refers to the components x and y that are responsible for most of
the effects attributed to the phenomenon of radioactivity.
6.1 Transformation processes overview 147
within one and the same nucleus K2, i.e. at both ΔZ and ΔA = 0 and may thus be
classified as “transitions” rather than transformations.
In addition, some of the primary and secondary transformation mechanisms, in
particular the electron capture process, leave a hole within the electron shell of the
nuclide involved. Once the new nucleus K2 has been formed, the vacancy in its elec-
tron shell must be filled. Finally, the particle emissions x (released in primary and
secondary transformations) interact with atoms surrounding the newly formed nu-
clide K2. The latter two independent effects are referred to as “post-effects”. Figure 6.2
illustrates the three process classes.
PRIMARY PROCESSES
ONLY NUCLEONS INVOLVED
SECONDARY PROCESSES
EXCITED NUCLEAR LEVELS
POST-PROCESSES I POST-PROCESSES II
CAUSED BY ELECTRON VACANCIES INDUCED BY EMITTED RADIATION
Fig. 6.2: Sketch of the interplay of primary transformations, secondary transitions and post-effects.
The transformation characteristics of a certain unstable nuclide, in particular the different types of
its “radiative” emissions, can only be explained as the sum of the corresponding processes.
PROCESSES ΔZ ΔA type
PRIMARY ≠0 -4
≈ -100 +
sf
≈ -140
SECONDARY =0 =0 , IC...
Fig. 6.3: Categorizing transformation routes of unstable nuclides ✶K1 ! K2 into primary and
secondary processes according to the balances of ΔZ and ΔA for the most dominant processes. The
corresponding subtypes differ in terms of the radiation co-emitted within the course of the
transformation and represent β-processes, α-emission or spontaneous fission, respectively. For
secondary processes, the most relevant case is emission of electromagnetic radiation (photons).
6.1.2.3 Post-processes
Electron capture means capturing an s-shell electron from the electron shell of K2
to allow the primary transformation of eq. (6.2f). After the new nucleus K2 has been
formed, the electron vacancy remains and must be filled by follow-up processes
(“post-processes I”). Independently, the components x released as particle emission
in eqs. (6.2c–e) interact with atoms surrounding the newly formed nuclide K2. The
effects induced by these interactions are discussed as “post-processes II”.
There is one exception, valid in a few cases of electron capture processes; this is shown in eq.
(6.1f) and Chapter 10.
150 6 Processes of transformations: Overview
different chemistry. Thus, chemistry is usually not important for nuclear processes,
but nuclear processes affect chemistry.
The corresponding subtypes of primary processes differ in terms of the way nucleon
compositions change within the nucleus of the atom. Consequently, unstable nuclides
convert into stable ones by either changing the absolute number of nucleons (chang-
ing A) or by modifying the ratio between protons and neutron (changing Z:N for con-
stant A), as per eqs. (6.2e, f). In the latter cases, an “excess” neutron “just” converts
into a proton (supposing the nucleus has an excess of neutrons over protons) or vice
versa. In other cases, the nucleus releases a number of nucleons, typically as a small
cluster of 2 neutrons and 2 protons – the α-particle – thereby lowering mass number
A (see eq. (6.2b)). For a limited number of very heavy nuclides, there is another option
to lower mass, namely a spontaneous split of the large nucleus into (mainly two) frac-
tions as per eq. (6.2g). This is called spontaneous fission (sf).
Interestingly, many unstable nuclides realize two or even more options; they sta-
bilize in parallel via α- and β-emission, or α-emission and sf, etc. This is because for a
given ✶K1, there is more than one option for fulfilling the ultimate prerequisite of ΔE > 0.
Table 6.1 illustrates the various transformation processes of unstable nuclides.
For a very limited number of extremely neutron-rich or proton-rich unstable nu-
clides, located close to the neutron drop line or proton drop line, a single “excess”
nucleon is emitted (eqs. (6.2c, d)).
However, for all naturally occurring unstable nuclides and most artificially pro-
duced unstable nuclides, β-emission, α-emission and sf are the most relevant pri-
mary transformation pathways. This textbook will concentrate on the β-processes,
α-emission5 and spontaneous fission.
The reason the β-processes are named “β” is historical and refers to the three sorts of emissions that
originate from naturally occurring unstable nuclides. These radiations were traditionally analyzed in
terms of charge and deviation in mostly magnetic fields, which work according to the ratio of q/mo.
This approach not only allowed to draw direct conclusions on the charge of the emission, but also to
derive its mass according to the deviation of the radiation. This is summarized in the well-known pic-
ture of the three emissions showing radiation emitted from unstable nuclides and analyzed in terms of
charge and deviation in electromagnetic fields. Following the Greek alphabet, the first radiation charac-
terized was the α-emission. The next generation of emissions of electrons analyzed this way was named
β. They were completely different to the α-emissions: charged negatively (−1), i.e. half of absolute
charge (α-particles are of + 2 charge), but of about 4000 times less mass. Consequently, deviation was
in the opposite direction, but significantly more pronounced. (The final emission categorized this way
was the γ-emission: no charge, no mass and thus not affected by the electromagnetic field at all.)
6.2 Primary transformation pathways 151
✶
Type Balance in mass Principal emission x Symbol K mass area
number ΔA
6.2.1.1 ΔA = 0
In most cases for a given nucleus, a proton just converts into a neutron or vice
versa. This type of transformation, the β-process, occurs throughout the Chart of
Nuclides, but is excusive for the light and medium heavy nuclides. In fact, β-
transformation processes are dominant for most unstable nuclides. More than 2000
nuclides are known to “utilize” this class of transformation. For nuclides of mass
number 2 ≤ A ≤ 144, it is the only possible class of transformation; the other two
main transformation types of ΔA ≠ 0, namely α-emission and spontaneous fission,
do not occur. For the ca. 3000 unstable nuclides, the proportion of nuclides prefer-
ring β-, α- or sf processes is roughly 100:10:1, as seen in Fig. 6.4.
6.2.1.2 ΔA ≠ 0
In the remaining cases, a nucleus releases one nucleon or (more likely) a small clus-
ter of 2 neutrons and 2 protons, the α-particle. This is a popular option for many
heavy unstable nuclides. For a number of very heavy nuclei, a third option appears,
namely to split the large nucleus into fractions. This only takes place above mass
number 232.
Figure 6.5 identifies the areas within the Chart of Nuclides, where the three
main primary transformation pathways dominate.
152 6 Processes of transformations: Overview
-TRANSFORMATIONS
Fig. 6.4: Frequency of the three primary transformation processes among naturally occurring and
most artificially produced unstable nuclides. The colors follow the color coding of the Karlsruhe
Chart of Nuclides. While yellow stands for α-emission and green for spontaneous fission, the
β-transformations are colored in either red or blue.
6.3 Energetics
114
a-EMISSION
SPONTANEOUS
FISSION
(184)
82
b+ EMISSION,
ELECTRON 126
CAPTURE
50
b- EMISSION
82
28
20 50
8
2 20 28
2 8
Fig. 6.5: Distribution of radionuclides in the Chart of Nuclides undergoing the β-processes, α-
emission and spontaneous fission. According to the Karlsruhe Chart of Nuclides, β-processes are
indicated by either blue (β−) or red (β+ or ε), α-emission by yellow, and spontaneous fission by
green.
and correlates with nucleon binding energy. The Q-value may also be calculated
using the mass defect of the components involved rather than the absolute masses.
However, something else is truly “emitted”. This other constituent is essential
for the impact of radioactive transformations in science and technology: the “radia-
tion” released by and accompanying the transformation processes. These emissions
are responsible for the effects generally associated with “radioactivity”. At this
stage, it is called “x” and/or “y”, and subsumes the various kinds of “radiation” to
be discussed later in detail.6
mnuclide
✶ K1 = mnuclide
✶ K2 + x ð + yÞ + Δ m (6:3)
Later, it will be specified that “radiation” consists of elementary particles (such as an electron),
subatomic particles (such as a positron and neutrino), or composites (such as an α-particle or neu-
tron(s)). Even later, in the context of secondary transformation processes, “x” will mainly denote
electromagnetic radiation such as photons and X-rays. For the latter, “X-ray” exactly reproduces
the historical context: emission of a kind of radiation not known that time (i.e. when ROENTGEN per-
formed his “X”-ray studies.)
154 6 Processes of transformations: Overview
6.3.1.1 Processes of ΔA = 0
Primary transformations that follow the pattern of eqs. (6.2e, f) are called “β-processes”.
These transformations proceed along an isobar line of the Chart of Nuclides, i.e.
with A = constant. For an unbalanced proportion of nucleons (unfavorable in the
context of mean nucleon binding energy) this may be addressed by either:
– substituting one proton with one neutron (if ✶K1 has an excess of protons over
neutrons) or
– substituting one neutron with one proton (if ✶K1 has an excess of neutrons over
protons).
The new coulomb term would be (Z ± 1)2/A⅓. The asymmetry term would count for
both nucleons as [(N − 1) − (Z + 1)]2/A or [(N + 1) − (Z − 1)]2/A (simplified to [(N − Z) − 2]2/
A and [(N − Z) + 2]2/A), respectively. Figure 6.6 illustrates the impact of these two terms
of the WEIZSÄCKER equation on the “improvement” of mean nucleon binding energy
along lines of constant mass number. The message is that internal changes of proton
to neutron (or vice versa) significantly affect the energetic – although the overall num-
ber of nucleons remains the same. Each of these conversions must yield a new nuclide
of (slightly) less absolute mass.
6.3.1.2 Processes of ΔA ≠ 0
Primary transformations following this pattern are the α-processes (eq. (6.2b)) and
spontaneous fission (eq. (6.2g)). For a given mass number A and nuclei with increas-
ing proton number, the coulomb repulsion becomes critical. Two terms of the WEIZ-
SÄCKER equation explicitly quantify the impact of the proton number directly, namely
the coulomb term Z2/A⅓ and the asymmetry term (N − Z)2/A. In the case of an α-
emission, the numerator Z2 turns into (Z − 2)2 and (N − Z)2 turns into ((N − 2) − (Z − 2))2,
despite changes in mass number from A to A − 4 for the denominator. This tendency
is reflected in Fig. 6.7, showing ĒB values depending on mass number A. For heavier
nuclei, each move towards lower A with ΔA = −4 slightly improves mean nucleon
binding energy.
6.3 Energetics 155
16
14
COULOMB + ASYM
COULOMB
10
3
ĒB (MeV)
8 4
0
0 50 100 150 200 250
A
Fig. 6.6: The principal effects of the various terms of the Weizsäcker equation on ĒB values versus
mass number A. These terms reflect the LDM of the nucleus of the atom: volume (1), surface (2),
coulomb (3), asymmetry (4) and pairing (5). Suppose that one sort of nucleon converts into the
other while mass number A remains the same; the coulomb term alone and the coulomb +
asymmetry terms together strongly modify the ĒB values.
An even more pronounced balance for ΔĒB values is achieved in the case of fis-
sioning a large nucleus. Nuclei of A > 230 have ĒB values less than 7.6 MeV, while
their fission products of mass numbers around ΔA ≈ −140 and ΔA ≈ −100 achieve
values of ĒB > 8 MeV.
156 6 Processes of transformations: Overview
9,0
9.0
8.5 -EMISSION
8,5
8.0 A-4
8,0
7.5
7,5
ĒB (MeV)
6.5
6,5
6.0
6,0
5.5
5,5
5.0
5,0
0 50 100 150 200 250
A
Fig. 6.7: The trend in mean nucleon binding energy according to the Weizsäcker equation with ĒB
values depending on mass number A for primary transformations of ΔA ≠ 0. For heavier nuclei,
each move towards lower A with ΔA = −4 slightly improves mean nucleon binding energy ΔĒB. For
very heavy nuclei, an even more extreme balance for ΔĒB values can be achieved by spontaneous
fission. Nuclei of A > 230 have ĒB values of less than 7.6 MeV, while their two fission products of
mass numbers around maxima at A ≈ 140 and A ≈ 100 (for e.g. 238U) show high mean nucleon
binding energies of ĒB > 8 MeV.
Let us consider a line of isotopes, as in Fig. 6.8. Suppose that the black square in-
volves a stable nucleon composition and isotopes to the right and left show increas-
ing or decreasing numbers of neutrons (at constant proton number). This will
(sooner or later) create proton to neutron ratios reducing overall and mean nucleon
binding energy as quantified by the WEIZSÄCKER equation, as per Figs. 4.1 and 4.3.
What happens next? Let us select the neutron-rich unstable isotope to the right of
to the stable nuclide. What should this A neutron-rich nuclide do? It may try to lower
the number of neutrons to get rid of the excess. Elimination of a neutron thus seems to
be a good idea. However, what does “excess” of neutrons mean? It is tantamount to a
6.4 β-Transformation processes (ΔA = 0) 157
“deficit” of protons. So, converting an excess neutron into a deficient proton would solve
the problem in a more elegant way. The same applies to the neutron deficient isotopes to
the left of the stable one, which are proton-rich. They could convert a proton into a neu-
tron. The point of the β-process is to convert the sort of nucleons that are in excess into
the other sort that are scarce. In doing so, the mass number will remain constant.
Conversion of a neutron to a proton follows the process ✶ z K1 ! Z + 1K2. This is ac-
companied by emission of a negatively charged electron and is called the β−-process.
Conversion of a proton to a neutron results in the opposite case, ✶ z K1 ! Z − 1K2.
n p
(b-)
p n
(b+, e)
Fig. 6.8: Schematic view on the origin and pathway of β-processes matching the Chart of Nuclides.
The middle horizontal line shows the isotopes of a given element. The black field in the middle
indicates stable nucleon composition, isotopes to the right (1) and left (2) show increasing or
decreasing number of neutrons (at constant proton number). These proton-to-neutron ratios tend
to reduce overall and mean nucleon binding energies as quantified by the Weizsäcker equation,
and (sooner or later) these nuclides are unstable. However, “excess” of neutrons is synonymous
with “deficit” of protons. Converting an excess neutron into a deficient proton (3) would thus
address the problem simultaneously. The same applies to neutron-deficient isotopes, which are
proton-rich, by converting a proton into a neutron (4). The mass number will always remain
constant. The successive steps of transformation thus proceed along the lines of isobars.
Conversion of a neutron into a proton at A = constant is the β−-process. Radioisotopes preferring
this pathway are color-coded in blue. Conversion of a proton into a neutron has two versions, both
color-coded in red: the β+-process and “electron capture”.
158 6 Processes of transformations: Overview
While there is only one approach for the ZK1 ! Z + 1K2 conversion, there are two
options for the ZK1 ! Z − 1K2 process. The one accompanied by emission of a posi-
tively charged electron, the β+ particle, is called the β+-process. Alternatively (or in
parallel), neutron-deficient nuclides may transform by capture of an electron from
the K electron shell. This type of β-process is aptly named “electron capture”.
ISOBAR LINE
p n
(b+, e)
ISOTOPE LINE
n p
(b-)
Fig. 6.9: β-Transformation of unstable nuclides. Neutron-rich nuclides transform via the neutron !
proton conversion (i.e. the β−-process). The new nuclide K2 has a composition of (Z + 1, N − 1) and
arrives at nuclides of higher Z (1). Proton-rich nuclides utilize proton ! neutron conversion (i.e. the
β+-process and electron capture) and yield a new nuclide K2 of (Z − 1, N + 1) composition (2).
Transformation may continue along the line of the same isobar stepwise, unless the Z to N ratio
represents a stable nuclide. It will have adequate mean nucleon binding energy according to the
liquid drop and shell models.
Fig. 6.10: β-transformation of unstable nuclides along the A = 18 isobar line with ĒB values in MeV.
The β+- and electron capture processes approach 18O from the proton-rich nuclides 18Ne and 18F,
while the β−-processes accumulate at 18O via 18B ! 18C ! 18N !.
There are many isobar lines across the Chart of Nuclides, ranging from short ones
(e.g. A = 3 with the two nuclides 3H and 3He) to very long ones (e.g. A = 100 includes
15 nuclides). An isobar line then represents directions of β−-transformation and β+-
and/or EC processes (see Fig. 6.11).
For β−-transformation, β+- and/or EC processes, there is ΔZ = +1 and ΔZ = −1, re-
spectively. The trend in ΔĒB = f(Z) at A = constant is shown in Fig. 6.12. The maximum
mean nucleon binding energy is located at the vertex of the parabola. Neighboring
nuclides are (in general) unstable. According to the systematic illustrated in Fig. 6.11,
mean nucleon binding energies increase more-or-less with every step (Z) approaching
stable nucleon configuration. Typically, the value of ΔĒB = f(Z) increases exponen-
tially and this is reflected by the exponential expression of a parabola, shown also in
p n
(b+, e)
n p
(b-)
A = CONSTANT
(ISOBAR)
Fig. 6.11: Unstable nuclides undergoing β-transformation processes along an isobar line of the
Chart of Nuclides.
Figs. 6.14 and 6.15. The blue nuclides from Fig. 6.11 (located on the right) shift to-
wards the left-hand side of the parabola in Fig. 6.12. This is because they are of low Z
compared to the red nuclides, which are of higher Z.
According to eq. (6.5), noninteger values may result for Z. In this case, the most stable nucleon
configuration is next to this value.
Equation (6.5) uses the five main terms of the WEIZSÄCKER equation. If the equation were modified
and extended, e.g. by the promiscuity term, this expression would slightly change.
6.4 β-Transformation processes (ΔA = 0) 161
NEUTRON
CONVERSION
PROTON
CONVERSION
ĒB
4 Z
Fig. 6.12: Unstable nuclides undergoing β-transformation processes as (1) β− route and (2) β+ or ε
route along a coordinate system of mean nucleon binding energy ĒB vs. Z at A = constant form a
parabola. Note that the y-axis showing ĒB is reversed – in contrast to how mean nucleon binding
energy correlates with mass number. (4). The maximum mean nucleon binding energy is located at
the vertex of the parabola (3). Interestingly (and unlike a normal parabola in algebra), the parabola
here gets two “ends”. The ends belong to the nucleon compositions that, according to the liquid
drop model, are energetically not permitted. They reflect the proton or neutron separation energies
forming the proton and neutron drop lines, as per Fig. 4.2.
A
ZA = 2
(6:5)
2.0 + 0.0154A =3
This is true for a single mass number A, and for the neighboring mass numbers as
well. Each isobar line in the Chart of Nuclides thus has a maximum of mean nu-
cleon binding energy for a specific value of Z.
Arranging these two-dimensional parabolas into a successive series creates a
three-dimensional plot, shown in Fig. 6.13. Now a two-dimensional parabola has
turned into a three-dimensional valley. Unstable nuclides are positioned along the
hillsides, stable nuclides at the bottom of the valley. The latter is called the “valley
of β-stability”. The direction of the valley does not correspond to a straight line
162 6 Processes of transformations: Overview
A = constant
ĒB A+2
A+1
A = constant
A-1
A-2
Fig. 6.13: The origin of the “valley of β-stability”. Individual parabolas (as shown in Fig. 6.12) are
arranged successively in order of increasing A, creating a “valley”. Stable nuclides are positioned
at the bottom of the valley, while unstable nuclides are along the hillsides.
(which would have been the isodiapher of (N − Z) = constant), but makes a sort of a
soft turn (curve) towards higher Z-values.
The line of the valley represents the positioning of the stable nuclides within
the Chart of Nuclides or vice versa. All the stable nuclides depicted in Fig. 3.3 of the
Chart of Nuclide diagram lie in this “valley”.
Characteristic for the mathematics of a parabola, the two ascents become steeper
and steeper. The x-axis, however, follows the linear scale of Z. This indicates that the
differences in mean nucleon energy between successive transformations of ✶K1 ! K2
are large at both “ends” of the parabola, and become less and less pronounced the
6.4 β-Transformation processes (ΔA = 0) 163
closer the transformation step is to the vertex. The sharper the ascent, the more un-
stable the nuclides. This is in good agreement with the LDM.
What about the complementary model of the nucleus of the atom: its shell
structure? Chapter 3 introduced the impact of specific occupancies of the various
nucleon shells. Ignoring magic numbers for a moment, even (= integer) nucleon
numbers are known to increase the mean nucleon binding energy. In part, this ef-
fect is covered by the WEIZSÄCKER equation too: it is the pairing term.10 What are the
consequences in respect to β-processes? Starting from a composition of ✶K1 = (Z, N),
each step of nucleon conversion creates a new nuclide K2 of either (Z − 1),(N + 1) or
(Z + 1),(N − 1). There are 4 different version of this process, as reported in Tab. 6.2.
Consequently, any change in ΔĒB caused by the coulomb term and the asymmetry
term is modified by the pairing term in a different way.
The impact of the pairing term is zero for (even, odd) and (odd, even) nuclides
✶
K1, because the new nuclide K2 is of the same category: (even, odd) turns into
(odd, even) and vice versa. This is the case for all isobars of odd mass numbers
A. Consequently, there is only one parabola for each mass number A, exemplified
in Fig. 6.14 for A = 95.
In contrast, an (odd, odd) nuclide ✶K1 turns into an (even, even) nuclide K2 and
an (even, even) ✶K1 turns into an (odd, odd) K2. In this case the pairing term mat-
ters. The change in ΔĒB caused by the coulomb term and the asymmetry term needs
correction as ΔĒB(COULOMB + ASYM) ± δLDM. This happens for all isobars of even mass
number A. It must yield two separate curves as indicated in Fig. 6.15 for mass num-
ber A = 96.
✶
K ! K Δδ LDM
The pairing term was introduced to reflect the positive impact of nucleon composition of (even
Z, even N) nuclei in terms of a coefficient δ of positive sign. In contrast, the negative impact of (odd
Z, odd N) nucleon mixtures was addressed by a negative sign for δLDM. For (even, odd) and (odd,
even) compositions, the coefficient is δLDM = 0.
164 6 Processes of transformations: Overview
ĒB
8.10 1
8.15 ĒB
3 A = 95
8.20 (o,e) (e,o) 2
8.25 3
8.30
ĒB (MeV)
2
8.35
8.40
8.45
8.50
8.55
4
8.60
5
8.65
8.70
35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50
ZA
Z
Fig. 6.14: β-parabola for mass number A = 95. Isobars of odd mass number A represent
transformations of (even, odd) nuclides ✶K1 into (odd, even) nuclides K2 and vice versa. In the
figure, (o,e) stands for (odd, even) and (e,o) for (even, odd). The impact of the pairing term is zero,
so there is only one parabola (1). (2) The successive transformations of type β+ or ε are shown at
the right-hand side, and of type β− at the left-hand side (not shown). (3) Decreasing differences in
mean nucleon binding energy are seen when approaching the vertex of the parabola. (4) ZA is
40.937. (5) The only stable nuclide is 95Mo (Z = 42) with the largest value of ĒB.
ĒB
8.10
1
8.15
A = 96
96Br
8.20 (e,e)
8.25
2 96Cd
8.30 3
ĒB (MeV)
96Kr
8.35
96A
8.40 96Rb
8.45 2
96Pd
8.50
96Sr
96Rh
8.55 5 4 5
96Y
96Tc
8.60 96Nb
96Ru
96Zr
8.65
96Mo 6
8.70
35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50
ZA
Z
Fig. 6.15: β-Parabolas for mass number A = 96. Isobars of even mass number A represent
transformations of (even, even) nuclides ✶K1 into (odd, odd) nuclides K2 and vice versa. In the
figure, (e,e) stands for (even,even). The impact of the parity term of the Weizsäcker equation
creates two parabolas, shifted in ĒB by ± δLDM. (1) Two separate parabolas arise; the one for (even,
even) nuclides is shown “below” (i.e., at higher values of ĒB). (2) The successive transformations of
type β+ or ε are shown on the right-hand side, and of type β− at the left-hand side (not shown).
These alternate from (even, even) to (odd, odd) nuclides and so on. (3) The shift in energy between
the two parabolas is Δ = 2δ. (4) ZA is 41.328. (5) The most stable nuclide is 96Mo (with the largest
nucleon binding energy). The stability of the two (odd, odd) nuclides needs to be studied in detail.
In this case, 96Ru is also stable, while 96Zr has a half-life of 3.9⋅1019 years (!)11.
This leads back to the question of the number of stable nuclides needed as an experimental
database for developing models on the structure of nuclei of atoms. Beta transformations create
stable nuclides for A = 1 – 209. Among these 210 mass numbers, there are 106 uneven numbers rep-
resenting (e, o) or (o, e) nuclides. In these cases, there is only one stable nuclide (except for Z = 61,
promethium). The other 104 mass numbers represent even values of A, and β-transformation switch
between (e, e) and (o, o) nuclei. Here, there is at least one stable nuclide of (e, e) composition repre-
senting the ultimate value of ĒB (except for Z = 42, technetium). This makes 105 + 103 = 208 stable
nuclides. In 42 cases, there are two stable (o, o) nuclei, and for mass numbers 124 and 132 there are
three, i.e. 208 + 42 + 2 × 2 = 254 stable nuclides. There are many (e, e) with very long half-lives, orders
of magnitude larger than the age of the earth. The question is, how many of the remaining (o, o)
166 6 Processes of transformations: Overview
Fig. 6.16: Excerpt of the Chart of Nuclides for nuclides of mass number A = 229, 231, 233, 235, 237,
239 and 241 with half-lives indicated: The need for a new primary transformation pathway. α-
emitting nuclides are color-coded in yellow. The dotted line represents the vertex of the β-
parabolas (1), the arrows indicate the two arms of the parabola, driven by β− (blue, (2)) and β+ and/
or EC (red, (3)) primary transformations. The β-transformation definitely ends at the vertex of each
parabola – but all the nuclides are (still) unstable (although of long half-life in the cases selected).
The α-emission proceeds at ΔZ = −2 and ΔA = −4 and is indicated for successive transformations of
241
Am ! 237Np ! 233Th (4). These transformations are parallel to the isodiaphere line (5). 233Pa
continues by β−-transformation (6) along its β-parabola of A = 233 = constant to 233U, which is at
the vertex of that β-parabola. It stabilizes by α-emission.
Still, there are isobar lines for mass numbers of A > 208. These reach values of
A = 252–262 for the heaviest lanthanide, lawrencium (Z =103) and approach 280 for
the transactinide elements. Figure 6.16 depicts the situation for unstable nuclides
of mass numbers A = 229, 231, 233, 235, 237, 239 and 241. The unstable nuclides rep-
resenting these isobars follow the β-transformation character as described. The indi-
vidual maximum of mean nucleon binding energy for those parabolas is reached for
nuclides are “really” stable? There are several candidates, such as 40K, 50V, 138La, 176Lu or 180Ta (t½
= 1.28⋅109 a, 1.4⋅1017 a, 1.05⋅1011 a, 3.8⋅1010 a or >1015 a, respectively), which are not “really” stable.
Only four are really stable: 2H, 6Li, 10Be, 14N. Thus, there are 258 “really” stable nuclides altogether.
6.5 α-Emission, cluster emission, spontaneous fission processes (ΔA ≠ 0) 167
the following nuclides: 229Th, 231Pa, 233U, 235U, 237Np, 239Pu, and 241Am. However, the
nuclides closest to the vertex of each parabola are not stable! And, there are no more
options according to the β-transformation profile!
This is the moment stabilization of unstable nuclides cannot continue following
the A = constant strategy. So, what should this unstable nuclide do? There are two
classes of transformations where A✶K2 < AK1: cluster emission (the most common is
emission of α-particles) and spontaneous fission (see Fig. 6.7).
6
3
Fig. 6.17: The relation between β-transformation and α-emission processes. α-emission occurs at
heavier mass number (1) and is most prominent at the vertex of isobar parabolas (2) of A > 209. It
either ends at a stable nuclide (3) or at neutron-rich parts of isobar parabolas of lower mass
number, and those unstable nuclides continue to stabilize via β-transformation within one more (4)
step along the corresponding isobar parabolas (5) to finally terminates at the stable nuclide (6),
either through a terminal α- or a β-transformation step (here shown for the latter version).
The α-emission process typically starts at the nuclide at the vertex of a parabola of
A > 209.12 Individual α-transformations or series of α-emissions proceed along a
straight line paralleling the isodiaphere. This either produces the stable nucleon
For unstable nuclides of increasing mass number of about A > 234, α-emission with
its mass difference of ΔA = −4 becomes a promising choice of transformation. How-
ever, there is another process that achieves even larger differences in ΔA, as shown
in Fig. 6.7. This occurs with very large nuclei. Figure 6.18 illustrates this option for
some of the nuclides already indicated in Fig. 6.16. Fission yields two fragments, K2
and K3, of the initial nuclide ✶K1 with characteristic mass distributions.
The fraction of nuclides utilizing spontaneous fission increases for very heavy
mass number, and for some of them this class of primary transformation becomes
the dominant pathway. Because very heavy nuclides have a significant excess of
neutrons over protons according to the liquid drop model (in particular represented
by the asymmetry term), the two lighter fragments have an even larger excess of
neutrons. Thus, they are unstable, and nuclide transformation continues along the
neutron-rich side of the β-process.
A representative example was mentioned in the context of the natural decay chains, as per
Fig. 5.23. The A = 4 n + 2 family includes the prominent nuclide 226Ra. This was generated from 234U
and 230Th via α-emission, continuing via α-emission to 222Rn, then 218Po and further to 214Pb. Here
the α-chain stops, and at A = 214 transformation is preferred following the β-characteristics “down-
hill” the neutron-rich side of the parabola at A = 214 = constant via 214Pb (β−)! 214Bi (β−)! 214Po.
The optimum value of ĒB(A =214) is reached here for unstable 214Po, and subsequent transformations
continue via α-emission.
6.6 Successive and simultaneous transformations 169
Fig. 6.18: Spontaneous fission for nuclides of mass number A = 231 to 241. Nuclides undergoing
fission are color-coded in green (circles (1)). One of the nuclides undergoing fission is 238U (2). It
splits into two fragments of statistic mass distribution, with one fragment peaking around A = 140
(3), the other around A = 100 (d).
However, a given unstable nuclide may not only “choose” one of the three primary
options – there are many radionuclides that can undergo two or even all three
routes of transformation. In this case, ΔE is positive for all pathways, but the indi-
vidual fractions are specific for each nuclide. For example, the prominent radionu-
clide uranium-235 shows a branch of α-emission and a branch of spontaneous
fission. Spontaneous fission occurs rarely, in just 0.0014% of the transmissions.
This means that given 1,000,000 atoms of 235U, the majority (999,986) transform
via α-emission, while only 14 undergo fission. The neutron-poor plutonium isotope
237 transforms basically via electron capture, but α-emission and spontaneous fis-
sion also occur. The neutron-rich plutonium isotope 241 transforms basically via β−
emission, but α-emission and spontaneous fission also occur, as shown in Fig. 6.20.
170 6 Processes of transformations: Overview
Fig. 6.19: Schematic comparison of the three main primary transformation processes of unstable
nuclei. The mass number A remains constant for β-transformation and proceeds along an isobar
line (1) and terminates at a stable nuclide positioned in the valley of β-stability (2). In the case of
an α-transformation, mass numbers A are successively reduced by 4 along an isodiaphere (3).
Because heavy nuclei start from a large excess of neutrons over protons (WEIZSÄCKER asymmetry
term), transformation product nuclides are even more neutron-rich. Consequently, α-emission
transits along β-transformation at the neutron-rich arm of the β-parabola. Thus, α-emission series
overlap with β-transformation. In the case of spontaneous fission, a nuclide of very large mass
number A splits into two fragments. It originates from a nuclide of very large excess of neutrons
over protons, and the two fission product nuclides are even more neutron-rich and quite distant
from the isodiaphere (4) of the initial nuclide. Each will undergo several steps of β-transformation
until a stable nuclide is formed.
A similar idea of co-existing transformation routes for one nuclide holds true
for the class of β-emissions. Nuclides such as 40K or 64Cu undergo all three sub-
types – though at different abundances – as per Fig. 6.21.
6.7 Secondary transitions 171
237Pu 241Pu
Fig. 6.20: Various transformation routes may apply to the same nuclide. The neutron-poor
plutonium isotope 237 transforms basically via electron capture, but α-emission and spontaneous
fission also occur. The neutron-rich plutonium isotope 241 transforms basically via β−-emission,
but α-emission and spontaneous fission also occur. The colors are red for β+ and EC, blue for β−,
yellow for α-emission, and green for spontaneous fission. The relative size of the different areas of
one nuclide square indicates the probability of a certain transformation pathway.
40 Ca 64 Zn
1 1
40 K 64 Cu
12.7 h
2 2
3
40 Ar 64 Ni
3
Fig. 6.21: Various transformation subtypes of β-emissions may apply to the same nuclide. 40K or
64
Cu undergo all three subtypes, though at different abundances. For 40K it is ≈89.3% β−-emission
(1), ≈0.001% β+-emission (2) and ≈10.7% electron capture (3), while for 64Cu it is ≈39.0% β−-
emission (1), ≈17.9% β+-emission (2), and ≈43.1% electron capture (3).
Chapter 3 introduced nuclear shell models applied to ground state levels of the nu-
clei. Primary transformation processes result in new numbers of protons and neu-
trons, i.e. the set of (Z, N) of the initial nuclide ✶K1 has changed.
However, not only the absolute number of each sort of nucleon matters due to the
LDM and the corresponding WEIZSÄCKER equation, but their arrangements do too. Pri-
mary nuclear transformation does not necessarily result in proton and neutron shell
occupancies that reflect the ground state levels of the corresponding nucleus as illus-
trated in Fig. 6.22. In contrast, the exothermic process of spontaneous primary trans-
formation seen in unstable nuclides allows for higher energy levels in the outer shells
172 6 Processes of transformations: Overview
of the nucleus to become populated. In fact, this is typically the case for all classes of
primary transformation: spontaneous fission, α-emission and β-nucleon conversion.
Direct, one-step processes from ✶K1 to the ground state of K2 are exceptional cases.
*K1 *K1
¤2 K2
¤1 K2
g K2 g K2
Fig. 6.22: Primary transformations (spontaneous fission, α-emission and β-nucleon conversion)
tend to populate one or more high energy nucleon shells in the newly formed nucleus ⊙K2, rather
than transforming directly and in one step to the ground state of K2.
There may even be several excited states ⊙iK2, depending on the number of sub-
shells available and the energy difference between them. With more protons and
neutrons per nucleus, the number of densely overlapping shell levels increases;
this is paralleled by an increase in potential excited states. Each nuclear state is de-
fined by the characteristic quantum parameters, such as nuclear spin and parity, as
discussed in Chapter 7. Rather than the excited state energies of nuclide K2, it is the
nuclear spin and parity of each nucleon which dictate why a primary transforma-
tion may terminate at a higher-energy nucleon shell. In other words, the unstable
nucleus K1 prefers an “easy” pathway, not a “hindered” one.
What could a “hindrance” look like? This refers to changes in spin and parity be-
tween the transforming nucleus K1 and the energetic state of K2. If both the spin and
parity values do not change, the transition is very easy – even called “super-allowed”,
cf. Figure 6.23. However, if the spin difference increases and the parity alternates, the
transformation is considered “forbidden”. This will be discussed further in Chapter 8.5.1.
Similar to electron excitation and de-excitation (i.e. relaxation) as discussed for
the hydrogen emission spectrum, excited nucleons tend to leave high-energy shell
6.7 Secondary transitions 173
2K2
t ≈ 0 or t½ of mK2
t ≈ 0 or t½ of mK2
1K2
t ≈ 0 or t½ of mK2
gK2
Fig. 6.23: Transitions between excited levels and between an excited and ground state. Transitions
either proceed very fast (within about 10−12 s) or the excited state remains for a much longer period
of time. This is indicated by having a “real” half-life.
levels for lower lying states in order to energetically relax. This corresponds to pro-
cesses between individual excited levels of type ⊙1K2 ! ⊙2K2. More generally, this
is written as ⊙iK2 ! ⊙(i − 1)K2, but can also appear as ⊙iK2 ! ⊙(i − 2)K2 and so on,
like the hydrogen electron transitions. In parallel, transitions can proceed from ex-
cited levels to the ground state, i.e. ⊙iK2 ! K2, which represents the lowest energy
available for the given nucleus.
This de-excitation stays within the nucleon configuration of K2, i.e. there are no
changes in proton and neutron numbers. The most frequently found mechanism for
this process is emission of energy in the form of electromagnetic radiation (γ-energy,
photons); see Chapter 11 for details of the secondary nuclear transformation processes.
The de-excitation processes in most cases proceed very quickly, with transitions being
terminated within about 10−12 s post primary nuclear transition. However, some ex-
cited states may “survive” for longer periods of time, as per Fig. 6.24. These states
thus exist for a much longer period of time than the usual transition speed. “Much”
longer is a relative term, and in this case indicates factors of 100 to 1000 or more. This
is known as “metastability”. By definition, metastability is the prolonged duration of
174 6 Processes of transformations: Overview
certain states14 and applies to several systems in physics and chemistry. It is also
known, for example, for electron excited states.15 The definition of metastable nuclides
actually refers – in analogy to chemistry – to isomer forms of the same nucleus. Those
species are thus also referred to as metastable isomers mK2, or simply “isomers”.16
energy
2 iK2
gK2 gK2
1
time
Fig. 6.24: Schematic illustration of metastable nuclei within successive transitions. (left) The
excited nucleus relaxes step-by =step, thereby minimizing the energy of the nucleus. The
dimension of the nucleus is depicted so that each step’s energy difference is too small for the
nucleus to remain there. It will immediately “fall down” the stairway to finally reach the energy of
the ground state (1) of the nuclide gK2. (right) One step (2) in the stairway is sufficiently wide for
the nucleus to remain for a longer duration in a metastable nucleon configuration mK2.
The excited level involved may be identified by its own half-life. The corresponding
state of ⊙K2 is called “metastable”. Metastable nuclide half-lives range from about
10−9 s to 1015 years (180mTa). The symbol of a metastable state of a nuclide becomes
m
K2, and in these cases the final ground state is denoted by gK. Within the isotope
notation, this gives 99mTc for the metastable technetium-99 (t½ = 6.0 h) versus 99gTc
for the ground-state technetium-99 (t½ = 2.1⋅105 a).
The Chart of Nuclides involves more than 3000 “boxes” identifying stable and un-
stable nucleon configurations. In addition, about 700 of these nuclides show one or
“Meta” refers to the virtual half-lives of the transitions between the other excited levels for the
same nucleus.
Excited electrons relax within about 10−8 s. Metastable electron excited states have durations of
about 10−3 s, i.e. several orders of magnitude higher. Within a cascade of subsequent electron tran-
sitions, this particular level will thus accumulate (or trap) electrons. This “population inversion” is
very important for the design and use of lasers.
In the early years of nuclear and radiochemistry, when for example the successive daughters of
the natural decay chains were analyzed chemically and spectroscopically, the nuclear parameters of
metastable nuclides were detected – but not necessarily attributed to an isomer of the same nuclide.
6.8 Post-processes 175
6.8 Post-processes
This isomer is used in nuclear medicine diagnosis by means of Single Photon Emission Com-
puted Tomography (SPECT). Due to its availability from a 99Mo/99mTc radionuclide generator and
the success of radiopharmaceutical chemistry to develop various biologically relevant 99mTc-
labeled molecules, 99mTc-SPECT imaging is performed in more than 100,000 patients worldwide
every day (discussed in Chapter 13 of Volume II).
Independently, the electrons of surrounding atoms represent a target for almost all kinds of ra-
diation emitted in transformations. When hitting an electron, radiation may release it and create a
so-called ion pair, consisting of the positively charged ion and negatively charged “free” electron.
Such processes represent the effect of radiation on condensed matter (in particular condensed bio-
logical “matter” such as humans) – and are the reason that radioactivity is commonly referred to as
“ionizing radiation”. They are not discussed in this textbook (Volume I: Introduction) because they
do not yield a new kind of “radiation”, but are covered in detail in Volume II: Modern Applications.
Chapter 1 is on Radiation Measurement (and obviously, this relates to the interaction of radiation
with well-defined condensed matter – the detector); Chapter 2 is on Radiation Dosimetry.
176 6 Processes of transformations: Overview
When traveling near the nucleus, beta-electrons of a certain initial kinetic energy can
be slowed down or deflected, reducing their kinetic energy. The difference in energy
is released as “bremsstrahlung”, and represents another, new kind of radiation.
6.9 Outlook
These aspects are covered in the following chapters for the three types of β-processes,
for α-emission, and for spontaneous fission.19 The following chapters focus on the de-
tails of secondary transformations and of the two post-processes.
In contrast, some very special transformations are not treated here, mainly because they are
almost irrelevant to nuclear and radiochemistry. One type is single or double proton emission,
which is observed for very proton-rich nuclides close or beyond the proton drip line.
7 β-Transformations I: Elementary particles
Aim: For most unstable nuclides the dominant pathway of transformation is the β-
process Its essence is turning either a neutron into a proton or vice versa, thus keeping
the overall mass number constant. It involves either emission or capture of an electron,
and there are three subtypes accordingly. These emissions and the secondary and
post-processes they induce all contribute to a complex spectrum of radiative emissions.
Nucleon conversion processes are a consequence of processes of elementary par-
ticles not yet discussed: quarks, electron neutrinos and anti-electron neutrinos, as
well as the corresponding mediators. Elementary particles are treated within the stan-
dard theory of particle physics. This chapter discusses the structure of quarks and
leptons (forming the group of fermions) and the field quanta (mediators) allowing
the interaction between fermions. These elementary particles form three families. In
addition, each fermion has its antiparticle, thus extending and completing the excit-
ing world of elementary particles.
Among the many different elementary particles, however, a rather limited num-
ber is needed in the context of β-transformations. These are the up quark and down
quark, the electron and the anti-electron (the positron), the electron neutrino and the
electron antineutrino, and finally the W bosons.
The characteristics of nucleon conversion also apply to nuclide transformations.
This process is defined by quantum physics in a theory called “FERMI’S golden rule”.
This is different to the other two primary transitions. When accepting a “liquid drop” model of a
nucleus of an atom, one would be willing to fancy the emission of an α-particle from a given nucleus,
or even the fission of a large nucleus into two fragments. In contrast, when thinking about “subatomic
particles” introduced so far, one should doubt the idea that one sort of a nucleon simply “turns” into
the other one.
https://doi.org/10.1515/9783110742725-007
178 7 β-Transformations I: Elementary particles
In order to understand the processes responsible for these events it is not sufficient
just to discuss the three subatomic particles introduced so far and their interaction –
instead, it is necessary to go beyond the structure of nucleons. While the three par-
ticles until this stage have been classified as “subatomic”, it does not necessarily
mean that all of them are not further divisible. This is true for the electron, which is
therefore classified as an “elementary particle”. This is not true for the proton and
the neutron! In contrast, the proton and the neutron are composed of other, subnu-
cleon particles. This is the moment to consult the concept of elementary particles.
Continuing in the tradition of the great Greek philosophers, elementary particles are
the ones that are not further divisible. Note that this allegorizes the ultimate idea:
Turning from visible matter to conceptual atoms, from experimentally proven atoms
to evidence of subatomic particles, and finally postulating the elementary particles!
According to the development of the standard theory of particle physics, there are
classes of elementary particles, arranged according to spin and electric charge. Spin
may be integer or half-integer, making “fermions”, and “bosons”. Fermions all have
spin ½. They involve elementary particles of integer electric charge (−1, 0, +1) and of
noninteger charge (−⅓, +⅔), termed “leptons” and “quarks” respectively.2
The fermions interact and the mediators3 allow for the interactions are called
“field quanta”. These are the photons, a group of “bosons” and a group of “gluons”.
Gluons mediate the strong interaction (strong in power, short in distance), attract nu-
cleons and are responsible for the formation of nuclei of atoms. The photon is the field
quantum mediating the electromagnetic interaction. Its range is longer, but the power
is just 0.1 relative to the power of the strong interaction. The W and Z bosons are corre-
lated with the weak interaction, of power just 10−12 compared to the strong interaction.
The elementary particles are summarized in Tab. 7.1. Each elementary particle is
unique (analogous to the PAULI principle) in terms of charge and spin. Among all the
elementary particles listed, the two most commonly known (at least in chemistry) are
the electron and the photon. The electron is of electric charge −1 and spin ½. It is
For example, the Nobel Prize was awarded to MURRAY GELL-MANN “for his contributions and dis-
coveries concerning the classification of elementary particles and their interactions”.
This reminds us of the concept of an atom as suggested about 24 centuries ago, when PLATO
(427–347 BC) added a fifth element, the ether, as the one allowing for interaction between the
others. According to ARISTOTLE (384–322 BC) these were the four components all existing matter is
built of (namely air, water, fire, and earth).
7.1 Elementary particles 179
thus a fermion and a lepton. The photon has no charge and its spin is 1. Accordingly,
it is a boson and belongs to the group of mediators. In Tab. 7.1 neither the neutron
nor the proton appears – neither are elementary particles.
Tab. 7.1: Overview on the system of elementary particles showing electric charge and spin.
Leptons electron e− – ½
e−-neutrino νe ½
muon μ− – ½
μ−-neutrino νμ ½
tau τ− – ½
τ−-neutrino ντ ½
Instead, there are quarks, defined by noninteger spin and also noninteger electric
charge (+⅔ or −⅓), and leptons, defined by half-integer spin and integer electric
charge (0 or −1). Both the six quarks and the six leptons belong to the class of fermions
(half-integer spin of ½). In contrast, the last group of elementary particles is the medi-
ators or field quanta, characterized by integer electric charge (0, +1, or −1) and integer
spin. Due to the value of spin = 1, this group belongs to the class of bosons.
Quarks and leptons are structured into three families, and – among other factors –
arranged according to their mass (or energy). Figure 7.1 provides a sketch of this
structure and of individual masses.
180 7 β-Transformations I: Elementary particles
Gauge Bosons
–1 –1 –1 ±1
½ e ½ μ ½ τ 1 W±
Leptons
mass
Fig. 7.1: Families of elementary particles and their field quanta. The first family includes the up
quark and the down quark, the electron and the electron neutrino. The second family collects the
charmed quark and strange quark, the muon and the muon neutrino etc.
7.1.3 Antimatter
The family of elementary particles thus lists 6 quarks, 6 leptons, and 12 bosons, repre-
senting altogether a world of 24 elementary particles. However, each elementary parti-
cle among the quarks and leptons has a “twin” that is identical in all parameters –
except the charge. Those are called “anti”-particles, shown in Fig. 7.2. This arises
from the criterion of symmetry in particle physics. Table 7.2 summarizes this for the
leptons and their antileptons. The most prominent antiparticle is the positron – at
least in the context of nuclear and radiochemistry. It has exactly all the properties of
the electron (mass, spin), but its charge is +1 instead of −1.
7.1 Elementary particles 181
MATTER
Leptons electron e− – ½
e−-neutrino νe
−
muon μ –
−
μ -neutrino νμ
−
tau τ –
τ−-neutrino ντ
Antileptons positron e+ + ½
+
e -neutrino νe
+
muon μ +
+
μ -neutrino νμ
tau τ+ +
τ+-neutrino ντ
182 7 β-Transformations I: Elementary particles
7.2 Quarks
Quarks4 are fermions due to their value of spin ½. They are the elementary particles
forming the matter of nucleons (and thereby the mass of nuclei and thus most of
the mass of an atom). Properties of quarks thus should reproduce the different
mass and charge of a proton and a neutron.
The group of quarks contains six different members, named “up”, “down”,
“charm”, “strange”, “top”, and “bottom”. Quarks have electric charges of multiples of
⅓ of an integer value and of either positive or negative sign. One group is of +⅔
charge, the other of –⅓. Up, charm and top quarks appear to be identical due to the
same spin and the same charge of +⅔ (but differ in mass), and the other three are
identical due to the same spin and the same charge of –⅓ (but differ in mass). Quarks
finally are understood to own a “color”, be red, green or blue5 – and thus each quark
is finally unique in terms of quantum number “philosophy”.
Although they differ in mass, the PAULI principle requires that every quantum
physical object has a set of quantum mechanics parameters that makes it unique.
This appears not to be the case. Quantum physics solves this problem by postulat-
ing one more property: the “flavor” of quarks, shown in Tab. 7.3. The combination
of these properties creates a difference among quarks of the same charge.
Tab. 7.3: Parameters describing quantum mechanics properties of quarks expressed as “flavor”.
I C S T B´
isospin charm strangeness topness (truth) bottomness (beauty)
u +⅔ +½
d −⅓ −½
c +⅔ +
s −⅓ −
t +⅔ +
b −⅓ + −
M GELL-MANN (1964) named quarks after JAMES Joyce’s Finnegan’s Wake: “Three quarks for mas-
ter Mark”. (R FEYNMAN suggested naming them “partons”.)
Like the theory on quarks overall, the definition of flavor and color are of course abstract.
7.2 Quarks 183
quarks electric
charge
u, c, t +⅔
d s, b
d, -⅓
The nucleons represent subatomic, but not elementary particles. The proton and the
neutron consist of quarks, see Figure 7.4. However, among the many quarks (six
“normal” quarks + six antiquarks) only two are relevant in the context of the β-
transformation processes: the up quark and the down quark, i.e. those of the first
family of elementary particles. The force the quarks interact in is the “weak interac-
tion”, mediated by bosons.
QUARK
NUCLEON NUCLEUS
Fig. 7.4: Three quarks make one nucleon. The nucleons build the nucleus of an atom.
If related to nucleons, the “magic” number for quarks is 3: Each quark has an electric
charge of a multiple of ⅓, each quark exists at 3 colors, and 3 quarks make one nu-
cleon. A proton is composed of two up quarks (2 times the electric charge of +⅔
makes a +4/3 charge) and one down quark (electric charge −⅓). The resulting total
charge thus is +3/3 = +1. For the neutron, there is one up quark and two down quarks
and their particular electric charges combine to the overall charge of 0 (see Tab. 7.4).
184 7 β-Transformations I: Elementary particles
Tab. 7.4: Composition of nucleons. The three quarks are responsible for
creating the different total electric charges of protons and neutrons.
Proton u + u + d +
+⅔ + +⅔ + −⅓
Neutron u + d + d
+⅔ + −⅓ + −⅓
d u
u d u d
NEUTRON PROTON
Fig. 7.5: Processes going on under the “surface” of nucleons, exemplified for the β−-process.
The metamorphosis of a quark elementary particle (here one d-quark into a u-quark) explains
the conversion of a neutron into a proton. The other down quark and the up quark of the
neutron remain unchanged. This explains the balance in quarks, but provokes another
question: What about the balance in charge? A neutral nucleon changes into a +1 charged
nucleon . . . so what?
– β+- and EC processes: For the conversion of a proton into a neutron, the inverse
occurs: a u-quark turns into a d-quark.
These are fundamental laws in chemistry. More than 200 years ago, LAVOISIER and PROUST stated the
law of conservation of mass and the law of definite proportions, respectively. In chemical reactions
and their stoichiometry, this refers to numbers of atoms (atoms cannot disappear or suddenly appear),
to the corresponding values of mole and mass, for example, but also for electric charge.
186 7 β-Transformations I: Elementary particles
The metamorphosis of one sort of first family member of quarks into another one
perfectly explains the balance in quarks, and perfectly explains the conversion of
one sort of nucleon into another one. However, several new questions arise. First,
what about the balance in charge? For β−-processes, a neutral nucleon had changed
into a +1 charged nucleon, while for β+- and EC processes a positively charged nu-
cleon had changed into a neutral one. As with conventional chemical reactions,
there must be a conservation of electric charge.
So, where is the missing charge going (for β−- and β+-processes) or coming from
(for the EC process)? To answer this, another elementary particle of the first family
is needed – the electron. Figure 7.6 illustrates how the electron takes care of a bal-
ance in charges within all the three subtypes of the β-process.
– β−-process: The emission of a “normal” electron within the n ! p conversion
satisfies the balance of electric charge: it is 0 ! (+1) + (−1).
– β+-process: The p ! n conversion requires the emission of the antimatter kind
of electron, the +1 charged positron.7 The balance of electric charge then is
(+1) ! (0) + (+1).
– EC process: The p ! n conversion can occur through another pathway, electron
capture (ε). Here, the proton captures a “normal” electron. The balance of electric
charge is then (+1) + (−1) ! (0).
Note that for all three subtypes of the β-transformation, an electron or positron is in-
volved. While the terminology “electron” is obvious in the case of electron capture, it
is more precise to call the electron released within the β−-process not an “electron”,
but a β−-particle. Similarly, the anti-electron ejected within the β+-transformation is
called β+-particle, or just positron. This is illustrated in Fig. 7.7.
7.3.5 Mass
Mass balance should consider masses of the individual components plus the value of
ΔE. Recall that energy is equivalent to mass and is much more relevant to nuclear
processes than conventional chemical reactions. Let us consider the case of the con-
version of a “free” neutron, i.e. one that is not part of an atom’s nucleus. The mass of
the neutron is a bit larger than the mass of the proton (1.675⋅10−24 vs. 1.673⋅10−24 g or
This creation of the positron is a daily practice in many of the facilities related to Positron Emis-
sion Tomography. Every day, positron-emitting radionuclides are produced and used for human
application at TBq radioactivity levels. Each transformation of a positron-emitting unstable radio-
nuclide thus produces antimatter. This application is part of Chapter 12 of Volume II.
7.3 Elementary particles relevant to β-transformations 187
+ e- +?
+ e+ +?
+ e- +?
Fig. 7.6: Balances in electric charge for nucleon conversion representing the three subtypes of
primary β-transformation. Gray and orange circles represent the neutron and the proton,
respectively. In all cases, it is an electron which handles the balance in charge, though in different
ways for the β−-process, the β+-process and the electron capture ε process. The question mark
indicates that despite the balance in electric charge, something more needs to be addressed!
e-
K-SHELL
Fig. 7.7: Different mechanisms for β−- and β+-transformations (left) compared to electron capture
(right). While for β−- and β+-transformations either a β−-particle or β+-particle is emitted, the ε
subtype needs interaction with a “real” electron from the shell of the atom ✶K1.
188 7 β-Transformations I: Elementary particles
1.00867 vs. 1.00728 u, respectively. For the unit energy see Tab. 7.5.).8 With the total
mass of the neutron, the proton and the electron known, the balance in mass for the
three subtypes of β-transformations can be derived, as shown in Tab. 7.5. The value of
Δm corresponds to the energy equivalent of the differences in mass ΔE = Q = Δm c2.
Tab. 7.5: Comparison of mass parameters involved in conversion of an unbound neutron into a
proton.
n ! p + e− + ΔE
Indeed, this conversion takes place for “free” neutrons (i.e. not bound in a nucleus
of a nuclide) with a half-life of t½ = 611 ± 1.0 s (= 10.18 min). In contrast, one may
believe that the conversion of a “free” proton into a neutron cannot occur, as mass
increases for the p ! n conversion. Actually, this is true for isolated protons; the
β+-process of unbound protons is impossible.
However, for bound nucleon systems, i.e. for nucleons located and interacting
within a nucleus of an atom, it becomes possible. In this case the “increase” in mass
for the p ! n conversion is “overcompensated” by the balance of the overall mass of
the two nuclides ✶K1 ! K2 involved. For example, the transformation of the proton-
rich and unstable isotope carbon-11 to stable boron-11 via the p ! n conversion (β+-
subtype) occurs according to the balance of the atomic masses of 11.011433 u !
11.009305 u respectively. There is a gain in mass of m(11C) – m(11B) = Δm = 0.002128 u,
despite the loss in mp – mn = Δm = –0.00139 u. Mean nucleon binding energy “im-
proves” according to 6.676 MeV ! 6.928 MeV for 11C ! 11B. In fact, the three pathways
of β+-processes are discussed in the following in the context of nuclide transformation,
i.e. refer to neutrons and protons bound in a nucleus of an atom.
While balances in electric charge and mass seem to be rather easy to address, this
is different with the momentum of the elementary particles and the nucleons in-
volved. Remember that quarks and electrons belong to the group of fermions – ele-
mentary particles of half-integer spin of value ½ each. The nucleons also show
values of spin ½. Conversion of spin for the β-processes now means: changes in
From this balance, it is obvious that a neutron may convert into a proton because the mass
balance mn – mp is positive. In terms of the atomic mass unit u, the difference is Δm = 0.00139 u.
7.3 Elementary particles relevant to β-transformations 189
spin between the initial and the final states of the transformations should be zero
or of integer value. So – here comes a problem!
Again let us consider the n ! p conversion of the free neutron. The neutron’s spin
is ½, so the total spin of the left-hand side of the transformation shown in Fig. 7.5 is
noninteger. Among the transformation products, the spin ½ of the proton and the
spin ½ of the electron combine to an integer number.
Fig. 7.8: The electron neutrino within the three subtypes of the β−-process. In all cases, an electron
neutrino ʋe is emitted in the context of nucleon conversions. It guarantees the conservation of spin.
This problem was addressed first by W PAULI, who in 1930 postulated the existence
of a hitherto unknown particle. It should have no electric charge (so he called it
“neutron”),9 (almost) no mass, both in order to not violate the balances in electric
charge and mass achieved for the β-transformation so far. However, it should carry
a half-integer spin. This electron neutrino then achieved the conversion of spin (see
Fig. 7.8). The neutrino hypothesis fits perfectly with all three subtypes of the β-process.
When J CHADWICK later (in 1932) discovered the “real” neutron, this became confusing. E FERMI
therefore introduced an alternative terminology by suggesting instead of “neutron” to use “neutrino”
(which is Italian and means “little neutral one”).
190 7 β-Transformations I: Elementary particles
H Cd
p n p
NEUTRINO
NUCLEAR REACTION
prompt
Fig. 7.9: Experimental proof for the existence of electron neutrinos10 via electron neutrino-induced
nuclear reactions and measurement of secondary processes.11 The electron antineutrino originated
from the transformation of a neutron-rich unstable nuclide, and it was supposed to interact
(although to a very low degree) with surrounding matter. This matter was hydrogen (as water),
providing protons in the form of its atomic nucleus. A neutron and a positron should appear as
reaction products (thereby fulfilling criteria listed above concerning the conservation of charge,
spin and symmetry). Detection was finally organized for the secondary processes, i.e. by
measuring the effects induced by the neutron and the positron that were formed. For the neutron,
this was a nuclear reaction with a nuclide with a high probability of capturing a neutron, e.g. 113Cd
(present in the water), undergoing a nuclear reaction leading to radioactive 114Cd. The latter was
radioactive, and when stabilizing it emits electromagnetic radiation (photons). The positron
underwent another process, which was combination with an electron (from the water molecule), to
convert the masses of the electron and the positron into two 511 keV photon quanta. These two
different photon radiations served as indicators that the process postulated in eq. (7.1) was true.
The neutrino first appeared only as a theoretical concept. The experimental detec-
tion of neutrinos, however, remained (and remains) extremely difficult due to the
extraordinarily small mass of this elementary particle. Historically, experimental
The other neutrinos (muon and tau) were detected much later, in 1962 and 2000, respectively.
The discovery of the muon neutrino was honoured by the Nobel Prize in Physics in 1988 “for the
neutrino beam method and the demonstration of the doublet structure of the leptons through the
discovery of the muon neutrino” to LEON M LEDERMAN, MELVIN SCHWARTZ and JACK STEINBERGER.
Characteristics of nuclear reactions such as thermal neutron capture and secondary processes
of nuclide transformations such as annihilation and photon emission will be discussed in following
chapters.
7.3 Elementary particles relevant to β-transformations 191
p + + ve ! n + e + (7:1)
7.3.8 Symmetry
All electron neutrinos should carry a spin of ½ to account for a noninteger spin of
the sum of the particles formed in β-transformations. In addition, it should exist as
a particle in one case (β+-process) and as an antiparticle in another case (β−-
process). Why? Whenever a new elementary particle is created, another comple-
mentary antiparticle should appear simultaneously, and vice versa.
In the case of n ! p conversion, the β−-electron formed is a “matter” particle. The elec-
tron neutrino should thus be an “anti-matter” particle, i.e. the electron antineutrino.
This holds true for the β+-process as well. Here, antimatter is created (the positron, the
β+-particle), thus a “normal matter” electron neutrino is needed. Figure 7.10 illustrates
the two faces of the electron neutrino, i.e. the electron neutrino and the electron anti-
neutrino, and lists the values of the relevant properties.
With electrons and electron neutrinos affiliated to the quark metamorphosis, all the laws
of conservation now apply and yield a more complete picture, summarized in Fig. 7.11.
It thus was an indirect approach that finally proved the existence of neutrinos. In 1959 COWAN
and REINES (Nobel Prize in Physics, REINES, 1995) investigated (i) the process of electron antineutri-
nos reacting with protons of the hydrogen nucleus, which was supposed to result (ii) in the forma-
tion of a neutron + positron pair, as per Fig. 7.9.
The experimental determination of the mass of the electron neutrino is an ongoing process. The
current data indicate mνe < 2 × 1.6⋅10−19 kg⋅m2⋅s−2 / c2 = 3.2⋅10−19 kg⋅m2⋅s−2 / (3⋅108 m⋅s−1)2 =
3.56⋅10−36 kg.
192 7 β-Transformations I: Elementary particles
_
-
e
d u
u d u d
NEUTRON PROTON
Fig. 7.11: Conservation of charge, mass, spin and elementary particle symmetry achieved
throughout the β-processes. Exemplified for the β−-subtype induced by converting a down quark of
a neutron into an up quark, creating a proton under simultaneous release of a β−-particle and an
antimatter particle: the electron antineutrino. But how is this transformation “mediated”?
Figure 7.12 illustrates the principal changes among the elementary particles in-
volved (up quark, down quark, electron or anti-electron, electron neutrino or elec-
tron antineutrino). Yet, there is one question: why should one sort of quark turn
into another one? A force is needed to manage this fundamental process.
The up quark, down quark, electron, anti-electron, electron neutrino and elec-
tron antineutrino all belong to the first family of elementary particles. This family is
the subject of the weak interaction and the field quanta responsible for carrying inter-
actions are the bosons. In the case of the β−-transformation, the boson in question is
the W− boson; for the β+-subtype it is the W+, and for electron capture the Z° boson.
FEYNMAN14 suggested graphical representations of this process (and many other pro-
cesses in elementary particle physics). This is shown in Fig. 7.12 with a focus on the
The Nobel Prize in Physics, 1965, was awarded to SIN-ITIRO TOMONAGA, JULIAN SCHWINGER, and
RICHARD P FEYNMAN “for their fundamental work in quantum electrodynamics, with deep-ploughing
consequences for the physics of elementary particles”.
7.4 Quantum theory of β-transformation phenomena 193
W-
d spectator u
quarks
u u
d d
Fig. 7.12: FEYNMAN diagrams of β−-transformation. Fermions are indicated by straight lines, bosons
by wavy lines, arrows indicate time directions. The down quark d (the “actor”) turns into an up
quark u. (The two other quarks are just “spectators”.) At the turning point, the field quantum
W− comes into action, indicated by a wavy line.
three quarks and the mediator, and in Fig. 7.13 with a focus on the role of the β-
particles / electrons and the electron neutrinos.
The resulting balances including the electron neutrino are listed in Tab. 7.6. For the
n ! p conversion, a β−-particle (rather than an ordinary electron) is emitted. It is a
particle and in this case, the co-emitted electron neutrino is an antiparticle (i.e., the
electron antineutrino). For the p ! n conversion, a positron is emitted. It is an anti-
particle (the anti-electron), so in this case the co-emitted neutrino is a normal ele-
mentary particle (i.e., the electron neutrino). For the electron capture process, the
electron neutrino is a normal elementary particle, too. Note that in this case no β-
particle is formed. All three nucleon conversions proceed inside the nucleus. For all
three versions an electron neutrino or antineutrino is ejected. However, β-particles
only appear for β−- and β+-transformations.
194 7 β-Transformations I: Elementary particles
_
e
-
e
+ e- e
W- W+ Zo
d u u d u d
Fig. 7.13: FEYNMAN diagrams of all subtypes of β-transformations. Fermions are indicated by straight
lines, bosons by wavy lines, arrows indicate time directions. In the β−-transformation for example,
the β−-particle emitted is shown with an arrow directed towards the right. An electron antineutrino
is co-emitted; its arrow points in the direction of the origin, which reflects the characteristics of an
“anti”-particle. β+- and ε-transformations: the up quark d turns into a down quark u. In case of the
β+-process, the field quantum W+ mediates the emission of the antimatter particle β+, and an
electron neutrino is co-emitted. In case of the electron capture subtype, the field quantum Z°
exclusively causes the emission of an electron neutrino.
-particle
f i
overlap
electron neutrino of wave
functions
Fig. 7.14: The phenomenon of β-transformation (left). Its principal concept in quantum theory
(right) showing wave functions Ψi and Ψf.
the two states. These states are expressed by a density, i.e. a number n of states per
unit of energy (dn/dE). The mathematics relates the probability (Pfi) of transition
(transition rate = transitions per unit time) to phase spaces via a matrix element
{Mfi}2. This matrix element considers the overlapping wave functions of the final
and initial states Ψf and Ψi and includes the Hamiltonian operator Ĥ of the weak
interaction. Suppose the overlap of the wave functions is large; then the probability
of transition is high.
2π ^ 2 dn
Pfi = ψ H int ψi (7:2)
h f dE
The idea of overlapping wave functions is illustrated in Fig. 7.15 for energies. The
energy level of the final state may be higher than that of the initial states. However,
transitions only occur to final states of lower energy.
Each state corresponds to a space phase volume with x, y and z coordinates
within a Cartesian coordinate system of pi. These are located on the surface of a
sphere as illustrated in Fig. 7.16. Each energy of the β-particle and the electron neu-
trino is described by its corresponding impulse coordinates pi and is:15 ΔxΔyΔz
ΔpxΔpyΔpz = ħ3. Energy and impulse of the β-particle correlate by E2 = p2c2 + m2c4.
For a given volume segment, only specific states are possible by V ≈ ΔpxΔpyΔpz. The
most relevant equations and their relationship are illustrated in Fig. 7.17. Equation
(7.6) finally relates quantum physics to energies of nuclide transformations.
In the case of the β−- and β+-subtypes of β-transformation, the nuclide transforma-
tion energy Qβ is in part delivered simultaneously to the β-particle and the electron
neutrino. These two elementary particles share the energy available. Consequently, β-
particle energies show a continuous spectrum, and the same is true for the electron
neutrino. The expression Qβ – Eβmax refers to the impact of the electron neutrino. The
i
Pfi f
Fig. 7.15: Overlapping wave functions for energy levels of initial and final states. Transition only
occurs to final states of lower energy with a probability of Pfi.
pz
py
Space phases
4 p2 dp / h3
px
Fig. 7.16: Space phase volumes within a Cartesian system of impulses pi. For a given volume
segment, specific states are characterized by ΔpxΔpyΔpz.
7.4 Quantum theory of β-transformation phenomena 197
4 4
dn = V p dp 2 dn = V p 2 dp
h3 h3
(7.3)
density dn / dQ
for Q = Ee + E as product dn = dne x dn
(b th are undependent)
(both d d t)
dn 16 2 dp
= V2 p 2p 2 dp
(7.4)
dQ h6 dQ
(
mass of and recoil of K2 negligible
E = Q – Ee = pnc
(7.5)
dn 16 2
= V2 p2 (Q – E max)2
dQ c3h6
d
dp
C = constant
Pfi(p)dp = n(p)dp
Mfi = transition matrix element
Pfi(p)dp = C Mfi
(7.6)
2 p2 (Q – E max)2 dp
Fig. 7.17: FERMI’s golden rule: Basic equations. Number of states for β-particle and electron
neutrino per volume segments (7.3) combine to densities of states for both elementary particles
(7.4). With negligible mass of the electron neutrino and very small recoil energy of K2, the density
of states is expressed in terms of overall energy Qβ of the transformation relative to the maximum
kinetic energy of the β-particle emitted (7.5). Probabilities of transition are finally described by the
transition matrix element Mfi, while several numerical parameters are combined to a constant
C = V2 / (2p3 c3 ħ7 ), eq. (7.6).
198 7 β-Transformations I: Elementary particles
Qβ values are not totally represented by the maximum value of the kinetic energy of
the β-particle; instead, a fraction of kinetic energy is left for the electron neutrino.
7.5 Outlook
The standard model of elementary particles is still under development. The current
status is to define quarks and leptons as fermions, and field quanta as b. Together
with the corresponding antiparticles, this already represents a large number of very
different particles. One essential elementary particle is the HIGGS, which was sought
eagerly for years and was discovered experimentally only in 2012.16 In addition,
there are new discussions towards an even richer world of elementary particles (!).
In nuclear and radiochemistry, the most relevant processes of nuclear transfor-
mations can fortunately be explained with a limited selection of elementary particles;
two quarks, the electron and its antiparticle, the photon, and the corresponding field
quanta. Though this set of elementary particles is essentially needed for understanding
the physical processes of β-transformations, the following chapters on the three main
primary transformations of unstable nuclides focus on the interaction of nucleons
(rather than sub-nucleon elementary particles). Nevertheless, the radiative emission
induced by those primary processes again refers to several elementary particles: β-
particles, electron neutrinos, and electromagnetic radiation.
FERMI’S theory can explain many important experimental findings relating to the
theory of the β-transformation. These include the maximum kinetic energy of β-
particles, energy distributions between β-particle and electron neutrino, and even cor-
relations between the energy of a β-transformation and its corresponding velocity
(transformations constant, half-life). However, as more facets of the β-transformations
appear, more theoretical concepts are needed. This is discussed in the following
chapter.
Nobel Prize in Physics, 2013, jointly awarded to F ENGLERT and PW HIGGS “for the theoretical dis-
covery of a mechanism that to our understanding of the origin of mass of subatomic particles, and
which recently was confirmed through the discovery of the predicted fundamental particles . . . ”.
8 β-Transformations II: β−-process, β+-process
and electron capture
Aim: For most unstable nuclides, the dominant pathway of transformation is the β-pro-
cess. Its essence is turning either a neutron into a proton or vice versa. It involves either
ejection of an electron from the nucleus (then called the β-particle) or capture of an
electron from the electron shell. The first subtype applies to neutron excess nuclides,
and because a negatively charged β−-particle is emitted, it is called the β−-process. For
neutron-deficient nuclides, β-transformations proceed either by emission of a posi-
tively charged β+-particle (the positron) and are called β+-processes. Alternatively or in
parallel, a proton of a neutron-deficient nucleus captures an electron from (mainly) the
K electron shell. This subtype is named electron capture (EC or ε).
This chapter discusses the individual characteristics of these subtypes. Topics
covered include how the overall energy ΔE is disseminated among the partners, in
particular in terms of kinetic energies of the β-particles emitted. The experimentally
obtained correlation between the gain in energy ΔE for the whole transformation and
the velocity of the transformation, i.e. its half-life, are presented. The theoretical and
mathematical models to describe these and more effects are introduced.
Finally, prominent β-emitters are mentioned, which are important to specific
aspects of nuclear and radiochemistry, to science and technology.
8.1 Phenomenon
The essence of the β-processes is turning either a neutron into a proton or vice versa.
It is the main primary transformation and characterized by ΔA = 0 and ΔZ = ± 1. As the
total number of nucleons does not change for the nuclides ✶K1 and K2, the transfor-
mations proceed along the isobars. Nucleon binding energies improve, which is best
expressed by the isobar parabola of ĒB = f(Z), shown in Fig. 8.1.
The first subtype applies to neutron-excess nuclides, and because a negatively
charged particle is emitted, it is called the β−-process. The second process, valid for
neutron-deficient nuclides, proceeds by emission of a positively charged electron and
is called the β+-process. Alternatively, or in parallel, a proton of a neutron-deficient nu-
cleus captures an electron from the K electron shell, and this subtype is named electron
capture (EC or ε). The β−- and β+-processes and electron capture show the β-particles
electrons released and the electron captured either on the side of the transformation
products or on the side of the initial nuclide, as shown in Fig. 8.2. The β−-process forms
a nuclide K2 of (Z + 1) composition and the new nuclide needs an additional electron.
In contrast, the β+-transformation forms a nuclide of (Z − 1) composition and one of the
initial electrons of the nuclide K1 becomes obsolete.
https://doi.org/10.1515/9783110742725-008
200 8 β-Transformations II: β−-process, β+-process and electron capture
p n
( +, )
n p
( -)
MeV)
3 2
ĒB (M
Fig. 8.1: β-Transformation along isobars and in terms of ĒB = f(Z). Neutron-rich nuclides transform
via the neutron ! proton conversion (i.e. the β−-process). The new nuclide K2 has a composition of
(Z + 1, N − 1). Proton-rich nuclides utilize proton ! neutron conversion (i.e. the β+-process and
electron capture) and yield a new nuclide K2 of (Z − 1, N + 1) composition. Transformation may
continue along the line of the same isobar stepwise, unless the Z to N ratio represents a stable
nuclide. A stable nuclide will have adequate mean nucleon binding energy according to the liquid
drop and nuclear shell models. Along the parabola of A = constant (1) one single parabola;
successive transformations of type β+ or ε at the right-hand side (2), and of type β− at the left-hand
side (3) improve mean nucleon binding energy and finally approach the vertex of the parabola
(4) with the largest value of ĒB.
While the main goal is to optimize nucleon binding energy among the unstable
nuclides involved, the β-processes are characterized by specific radioactive emis-
sions. Electron neutrinos are released during the primary transformation steps of
all subtypes. More relevant to nuclear and radiochemistry is the emission of the two
kinds of β-particles within the β−- and β+-processes.
8.2 Energetics of β-transformations 201
+ -
-PARTTICLES
RELEAASED
+
A
CAPTTURED
ELECTTRON
+ e-
Fig. 8.2: Release of the β-particles and capture of the electron involved in the three
β-transformation subtypes relative to the balance of the nucleons.
Tab. 8.1: Important nuclides ✶K1 undergoing β-transformation, their transformation product nuclide
K2, the Qβ-values and the half-lives. (May I suggest indicating what a, m, d etc represent in the
caption? Units of t1/2).
✶ ✶
K K Qβ (keV) t½ K K Qβ (keV) t½
−
β β +
H He . . a C B . . m
C N . a N C . . m
Na Mg . . h O N . . m
P S . . d F O . . m
S Cl . . d Na Ne . . a
Co Ni . . a Ga Zn . . m
Sr Y . . d I Te . . d
Y Zr . . h ε
Mo Tc . . h Ge Ga . d
I Xe . . d I Te . . d
Cs Ba . . a Er Ho . . h
Lu Hf . . d Tl Hg . . h
Use of mass Δmexcess: For the β−-transformation 131I ! 131Xe, for example, this would give Qβ = Δmexcess
(131Xe) − Δmexcess (131I) = 88.414 MeV − 87.443 MeV = + 0.9701 MeV. For the β+-transformation 18F ! 18O,
for example, it is Qβ = Δmexcess (18O) – Δmexcess (18F) = (− 0.783 MeV) − 0.873 MeV = + 0.0903 MeV.
8.2 Energetics of β-transformations 203
The way the Q-value is calculated, however, is in part modified according to the
role of the electron captured and the β-particles emitted. The β−- and β+-particles do
not belong to the mass m of the nuclide ✶K1 nor to K2. In contrast, the electron cap-
ture subtype starts at a state of nuclide ✶K1, collecting one additional electron on
top of the initial electron shell configuration of the corresponding atom.
A*K1 AK +
Z K N Z-1K2 N+1 + + e
A*K1 + e- A*
Z N Z-1 K2 N+1 + e
Fig. 8.3: Mass balances for β+- and ε-transformation identifying the different impact of the positron
emitted and the electron captured. For both subtypes, a proton-rich nuclide ✶K1 forms a nuclide K2
of (Z − 1) composition. The difference is that for β+, one more electron mass appears at the side of
the products (which is the positron emitted). Likewise for ε, the mass of one electron captured
counts on the other side of the balance, i.e. its mass must be added to that of nuclide ✶K1.
Consequently, the balance in electron rest mass is either 2 or 0.
Using the masses m of the nucleus and electrons me (calculated as Z times the rest
mass of the electron me), values of Δm are calculated as illustrated in Fig. 8.3. Re-
gardless of the difference in mass between the two nuclides, the β+-transformation
requires an excess of that Δm plus 2⋅me. The amount of energy which equals the
mass of two electrons is 2⋅me⋅c2 = 2⋅0.511 MeV = 1.022 MeV. In contrast, electron
204 8 β-Transformations II: β−-process, β+-process and electron capture
capture (and β−-processes) are energetically satisfied by mK2 < m✶K1. This in particu-
lar discriminates between pathways of proton-rich unstable nuclides, i.e. β+- and ε-
transformation.2
The electron shell is not involved at all in the β−- and β+-transformations. In con-
trast, the ε-transformation needs an electron to be captured by one proton of the
nucleus of an unstable atom. In Chapter 1 the structure of the atom was introduced,
stating that the diameter of the nucleus is smaller than the whole atom by a factor
of about 1000. Thus, the distance between nucleus and shell is enormous in the
context of dimensions of nucleons or even electrons.
Nevertheless, an electron can be captured by the nucleus ✶K1 from its surround-
ing electron shell! Chapter 1 introduced the BOHR model, which assigns electrons
into separate shells with different energies. Quantum physics identifies each elec-
tron by a characteristic set of quantum numbers. And here comes a very specific
feature of electrons of orbital quantum number l = 0, the s-electrons: Because of
their spherical orbital distribution, they have – unlike all the other l > 0 shells – a
nonzero probability of existing close to (or even inside!) the nucleus.3 Figure 8.4 il-
lustrates that as shells of higher n are filled, lower-energy s-electrons (shells of
lower n) are more attracted by (and get closer to) the nucleus.
1s 2s 3s
2s
1s 1s
Fig. 8.4: Changing dimensions of s-electron orbitals for main quantum numbers n = 1, 2, 3.
Example: proton-rich 7Be (t½ = 53.29 d) transforms to stable 7Li. Atomic masses are 7.016003 u
and 7.016929 u, respectively; thus Δm is 0.000925 u = 0.863 MeV. (Utilizing mass excess values,
this is Qβ = Δmexcess (7Be) − Δmexcess (7Li) = 15.769 MeV − 14.907 MeV = + 0.862 MeV.) Consequently,
the only pathway open for 7Be is electron capture.
Note that this generally refers to the particle–wave duality of quantum physics.
8.2 Energetics of β-transformations 205
In the context of β-processes, proton-rich unstable nuclides have two options to im-
prove their nucleon binding energy: β+-emission and electron capture. The impact
of energy was discussed above. The first criterion is energy in terms of mass balan-
ces. Electron capture needs a mass balance of m✶K1 > m✶K2, while the β+-option ap-
pears only at m✶K1 > m✶K2 + 2 me (see Fig. 8.3). At ΔE > 1.022 MeV; the question
remains, why should a given proton-rich unstable nuclide prefer one pathway to
the other? A further consideration is how the ratio between electron capture and
the β+-process for similar proton number Z depends on the Q-value of the two sub-
types. In this case, the β+-process increases in relative frequency with increasing
Q-values.
Finally, the spatial distribution of electrons is defined by the orbital quantum
numbers l, giving rise to s, p, d, f (etc) orbitals. As mentioned, the spherical three-
dimensional distribution of the s-orbital overlaps in part with the nucleus of the
atom. Thus, there is a quantum mechanics probability (albeit very low) for s-orbital
electrons to exist “inside” the nucleus. On a relative scale, this probability is most
pronounced for K-shell electrons over L- or M-shell s-electrons. Electrons with n > 1
have higher energies and orbit at greater distances from the nucleus (see Fig. 8.4).
From this, several conclusions can be drawn:4
1. The probability of electron capture increases with decreasing distance of the K-
shell to the nucleus. The distance between the nucleus and K-shell follows a
function of 1/Z2, as per Chapter 1. The higher the element’s proton number Z,
the higher the probability of electron capture.
2. Consequently, electron capture dominates in the case of unstable proton-rich nu-
clides of heavy elements. Likewise, β+-emission dominates for light elements.
The corresponding ratio between electron capture and the β+-process thus depend on
the proton number Z of ✶K1. A ratio of 1 indicates that both pathways are equally prob-
able. In fact, most of the proton-rich unstable nuclides proceed via both pathways,
though at different percentages. For medium and large Z nuclides, K-shell s-electron
Theoretically, there may even be a third conclusion: for a chemical element (in particular of high
Z number), the chemical oxidation state may influence the quality of electron capture. For the ele-
ment technetium, for example, oxidation states range from + VII (TcO4−, pertechnetate) to –I (car-
bonyl complexes), revealing almost all the intermediate oxidation states for individual chemical
species. The distance of each electron shell (not just the valence shell) from the nucleus depends
on the ratio between the positive and negative charge carriers (i.e., protons and electrons). With a
decreasing number of outer electrons (oxidation state + VII ! –I) the electron shell dimensions ex-
pand. Consequently, the probability of electron capture should decrease for a number of e.g. 99mTc
species of oxidation state + VII ! –I. This may – at least theoretically – even correlate with changes
in the physical half-life of one and the same radionuclide. This would represent an exception from
the general statement that processes of radionuclide transformation do not depend on chemistry.
206 8 β-Transformations II: β−-process, β+-process and electron capture
capture starts to dominate. In contrast, the β+-process dominates for proton-rich un-
stable nuclides of relatively low Z.
Unstable proton-rich nuclides which preferentially undergo β+-transformation are
found among the second period of the Periodic Table of Elements. These include car-
bon (11C, t½ = 20.38 min), nitrogen (13N, t½ = 9.96 min), oxygen (15O, t½ = 2.03 min),
and fluorine (18F, t½ = 109.7 min). Abundances of the β+-subtype are 99.76 %, 99 %,
99.9 % and 96.7 % for 11C, 13N, 15O and 18F respectively. These are key nuclides in med-
ically important molecular imaging and diagnostics via positron emission tomography
(PET) and find extensive application in radiopharmaceutical chemistry. Nevertheless,
some unstable nuclides of elements above Z = 20 also emit positrons at percen-
tages which are relevant for practical applications. However, the emission per-
centage drops with increasing proton number Z: 68 Ga = 88.0 %, 73 Se = 65.0 %,
86
Y = 34.0 %, 89Zr = 23.0 %, 90Nb = 51.1 %, 124I = 24.0 %, for example.
Kinetic energy and momentum are relevant properties of all the species formed in
the primary nuclear transition, i.e. to the new nuclide K2, the β-particle and/or
the electron neutrino emitted. For example, the transformation products for the
β−-transformation are K2 + β− + ν̅e. The overall energy ΔE = Qβ − is shared among
all the products. The question is: how?
Momentum5 p is m⋅v (mass m, velocity v) and kinetic energy is E = ½m⋅v2. Mo-
mentum is conserved for all the species via mi⋅vi = mj⋅vj. The mass distributions
are very different for the three species formed. The rest mass for an electron and
electron neutrino are me = 9.1094⋅10−31 kg and 3.56⋅10−36 kg, respectively. The ab-
solute masses m of K2 depend on the mass number of the nuclide. For example,
this mass can be 10 u (for A = 10) or 100 u (for A = 100). In terms of absolute mass,
according to m = A⋅u, these values are 1.66054⋅10−25 kg or 1.66054⋅10−24 kg, re-
spectively. The nuclide masses are thus about six orders of magnitude larger than
the electron rest mass.
Conservation laws refer to momenta of K2, the β-particle and the electron neutrino.
The overall energy Q is disseminated between the species according to eqs. (8.7–8.9).
To discriminate from the symbol p (used for the proton), momenta are denoted by an italic p.
8.3 Kinetic energetics of β-transformation products 207
The higher the Q-value, the higher the maximum kinetic energy of the elementary
particles emitted.
Because mass values are very low for the β-particle and the electron neutrino
compared to K2, their kinetic energies are opposite, as mβ⋅Eβ ≈ mK2⋅EK2. In the case
of β−- and β+-transformations, both elementary particles share this main fraction of
energy Qβ, which is less than the total energy of the nuclear transformation (see
eqs. (8.7) and (8.8) for the care of β− transformation). This is different for electron
capture, because no β-particle is emitted. In this case, the kinetic energy is shared
between K2 and the electron neutrino only, following eq. (8.9).
Eν e = Qε − EK2 (8:9)
Momentum and kinetic energy, however, must consider the high speed of the low
mass β-particle and neutrino. The β-particle and the electron neutrino have relatively
high velocities and thus differences between rest mass mβo , i.e. at zero kinetic energy,
and “real” mass mβ of the β-particle, at any velocity relatively close to the speed of
light c, must be considered. For relativistic velocity, eqs. (8.10) and (8.11) apply.
2
E2β = p2 c2 + m0β c2 (8:10)
Eβ = mβ − m0β c2 (8:11)
Let us assume the β-particle is ejected from K2, i.e. the former ✶K1. The impulse of
the β-particle causes a somehow opposite impulse to K2. This is referred to as “re-
coil energy” of K2. It is linked to:
(a) the Q-value,
(b) the mass of K2,
(c) the kinetic energy Eβ of the emitted β-particle and the electron neutrino6 (or the
electron neutrino exclusively in case of electron capture).
In addition, recoil energy is influenced by the spatial arrangement of the two elemen-
tary particles during emission (see Fig. 8.5).The K2 recoil energy thus lies between
The impact of the electron neutrino on the recoil energy can be neglected due to the extremely
low mass of the electron neutrino.
208 8 β-Transformations II: β−-process, β+-process and electron capture
RECOIL max
the theoretical maximum value and zero. The maximum kinetic energy EK2 a
recoil nucleus may receive is calculated with eq. (8.12).
! max
Emax
β Eβ
EK2 =
RECOIL max
2
+ mβ0
(8:12)
2c mK2
For example, the β−-transformation of the carbon radioisotope 14C into 14N yields
RECOIL max
EK2 = Emax (14N) = 6.9 eV. (mK2 = 14 u, m0β =0.511 keV, Emax
β = 0.156 MeV).
-particle
K2
recoil
electron neutrino
Fig. 8.5: Simplified two-dimensional illustration of spatial emission profiles of a β-particle and an
electron neutrino that influence the recoil energy of K2. Absolute recoil energy depends on the
mass of K2, the kinetic energy of the emitted β-particle and/or the electron neutrino and the
angular distribution the two elementary particles during emission.
The recoil energy of K2 is higher when its mass number is low and the emitted
β-particle’s kinetic energy is high. For example, at mass number A ≈ 100 and a maxi-
mum kinetic energy for the emitted β-particle of 1 MeV, the value of RECOILEmaxK2 is
about 10 eV. Table 8.2 gives typical values for other mass numbers and β-particle
7
energies.
Recoil energies are, in many cases, below the electron binding energies defining chemical bonds
between atoms, discussed in Chapter 1. However, in some cases, the energy is above that order of
magnitude and thus may break chemical bonds. This is a special topic in nuclear and radiochemistry.
8.3 Kinetic energetics of β-transformation products 209
RECOIL max
EK2 (eV)
Emax
β (MeV) . . .
AK =
AK =
AK =
Since the recoil nucleus receives a very small amount of kinetic energy, the dominant
fraction is left for the small particles emitted. For electron capture, all the remaining
kinetic energy goes to the electron neutrino, as per eq. (8.9). Consequently, the elec-
tron neutrino receives a kinetic energy of a discrete energy value.
However, for the β−- and β+-transformations, this is different, as per eq. (8.8).
How do the β-particles and the electron neutrinos share their fraction of kinetic en-
ergy? The answer is: “statistically”. There are cases where the β-particle gets all the
kinetic energy and nothing is left for the electron neutrino – or vice versa. In the first
case, the kinetic energy of the β-particle is at maximum and denoted by Emax β . Its
value depends on the value of Qβ. In reality, there is a distribution between both the
elementary particles, and consequently, kinetic energies observed for β-particles and
for electron neutrinos show a continuous spectrum. The β-particle kinetic energy
thus lies between the theoretical maximum value and zero. The typical maximum en-
ergy for β−- and β+-particles emitted from neutron-rich and neutron-poor unstable nu-
clides, respectively, ranges from about 20 keV to a few MeV.
Figure 8.6 shows a simplified course of kinetic energy for β−-particles emitted
from unstable 3H and 14C. The y-axis shows the relative number of β-particles ob-
served for a given energy. Maximum kinetic energies Emax β − are 18.591 keV and
156.476 keV, respectively, but the fraction of β-particles that reaches this energy is
very low. Only a small fraction of the β−-particles shows this maximum energy or
energies close to this value. More than ⅔ of the β-particles emitted have energies
less than ⅔Emaxβ− .
Most β−-particles typically show an average or mean energy (Emean β . or Ēβ)
around ⅓Emax β . However, both the very low energy and maximum kinetic energy of
β-particles requires a deeper look (see Fig. 8.8).
The same applies to positrons emitted within the β+-subtype of β-transformation.
Figure 8.7 shows profiles of the continuous spectra of the positrons emitted from four
relevant nuclides used in medical diagnosis (PET).
210 8 β-Transformations II: β−-process, β+-process and electron capture
3He 14N
Number of particless (normalized)
3H 14C
emitted per energy (dN( / dE))
12.323 a 5730 a
0.02 0.2
kinetic energy E
Emax (3H) E max(14C)
(keV)
0.0030
0.0025 18F
particless emitted pe
0.0020
11C
(d / dE)
0 0015
0.0015
13N
0.0010
-
15O
Number of
0.0005
0
0 400 800 1200 1600
E + (keV)
Fig. 8.7: Theoretical β+-particle energy spectra of 11C, 13N, 15O and 18F. Spectra are normalized to
have the same area under the curve. Maximum values of kinetic energy differ depending on the
Q-values, as per Tab. 8.1. With permission: CS Levin, EJ Hoffman, Calculation of positron range
and its effect on the fundamental limit of positron emission tomography system spatial resolution,
Phys. Med. Biol. 44 (1999) 781–799.
8.3 Kinetic energetics of β-transformation products 211
particles
-
relative number of
E max
E
IMPACT OF COULOMB FORCES IMPACT OF MASS
INDUCED BY THE NUCLEUS OF THE ELECTRON NEUTRINO
Fig. 8.8: Continuous spectrum of β-particle kinetic energy and areas needing additional comments.
This discussion refers to the low kinetic energy part of the theoretically continuous
β-particle energy spectrum. The abundance of β−-particles sharing their kinetic energy
with the electron neutrinos is explained by eq. (8.8). However, there is a special effect
caused by the different coulomb interaction between the positively charged nucleus
of the radionuclide and the negatively or positively charged β−- and β+-particles emit-
ted, summarized in Fig. 8.9. It seems obvious that the negatively charged β−-particles
“feel” some coulomb attraction caused by the nucleus – they are therefore a bit
“slower” than theoretically predicted. In other words, the abundance of low-energy
β−-electrons observed is higher than the theoretical line of the continuous energy
spectrum of e.g. Fig. 8.8. In contrast, positively charged β+-particles (positrons) re-
ceive some coulomb repulsion, which increases their initial kinetic energy. Extremely
speaking, there are almost no positrons at zero or close to zero theoretical kinetic en-
ergy, because they have obtained an additional fraction of kinetic energy through
coulomb repulsion.
Quantum physics treats this phenomenon mathematically. It introduces a so-
called Fermi correction term F(Z,Eβ) into the basic equation (7.6), shown in Fig. 8.10.
It may be understood as a numerical factor being either >1 or <1, respectively, for
β−- and β+-particles.
212 8 β-Transformations II: β−-process, β+-process and electron capture
- particle
p
+ particle
Fig. 8.9: β-particle attraction (β−) or repulsion (β+) influencing the energy distribution of the
low-energy particles emitted in the course of β−- or β+-transformations.
C = constant
Mfi = transition matrix element
(7.6)
Pfi(p)dp = C Mfi 2 p2 (Q – E )2 dp
COULOMB interaction
between electric charges of nucleus and ±-particle
(in particular at large Z and low E )
F(Z,E ) > 1
F(Z,E ) < 1
(8.13)
Fig. 8.10: Fermi correction function. The term F(Z,Eβ) “corrects” for the coulomb interaction
between the positively charged nucleus and the negatively charged β−-particle or the positively
charge β+-particle. The term F thus depends on the proton number Z (if high) of the initial unstable
nuclide and the initial kinetic energy of the β-particle (if low), as per eq. (8.13).
Because the charge (+1 or −1) and mass of β-particles are always the same, this
refers (a) to the small or large number of protons in the nucleus (Z), and (b) to the
initial kinetic energy the β-particles obtain according to the Q-value. As a consequence
of the latter effect, it will only modify the “early” part of the continuous β-particle
spectrum, i.e. the lowest-energy β-particles.
8.3 Kinetic energetics of β-transformation products 213
This discussion refers to the high kinetic energy part of the theoretically continuous
β-particle energy spectra, shown in Fig. 8.8. In the case of kinetic energies close to
the maximum value Emax β , the kinetic energy of the co-emitted electron neutrino (at
least theoretically) modifies the spectrum. The very low, but definitely >0 mass of
the electron neutrino must be subtracted from the measured endpoint energy of the
β-particle. Figure 8.11 shows this effect for the spectrum of the β−-transformation of
tritium.
– neutrino mass
4 10–8
relative number of - particles
1.0 eV
0.3 eV
0.0 eV
3 10–8
E max
-
-
2 10–8
ΔE = 1.0 eV
1 10–8
0
18.598 18.599 18.600
E - (keV)
Fig. 8.11: Part of the energy spectrum of the β−-transformation of tritium lying very close to the
maximum energy values. Detected within the Katrin experiment to determine the mass of the
electron neutrino. Spectrum endpoint energies are corrected for three possible masses of the
electron antineutrino: 0 eV, 0.3 eV and 1.0 eV. (https://de.wikipedia.org/wiki/Katrin).
illustrated in Fig. 8.12, otherwise known as a KURIE plot.8 It is mainly relevant in de-
riving experimental values for Emaxβ from experimentally measured continuous spec-
tra of β-particles.9
(8.14)
n(p)dp = C Mfi 2 F(Z,E ) p2 (Q – E max)2 dp
n(p)dp
(8.15)
___________ (Q – E max)2
F(Z,E ) p2 dp
_________
n(p)
(8.16)
_______ max
Q –E
F(Z E ) p2
F(Z,E
Fig. 8.12: Kurie plot: Quantum physics correlations reflecting “pure” energy values Qβ and Eβ in a
linear relationship.
Deviations from linearity close to endpoint energies are of particular interest; they
may indicate the impact of the electron neutrino as illustrated in Fig. 8.11. KURIE
plots are thus a tool to determine parameters of electron neutrinos. Figure 8.13 de-
picts how the mass of the electron neutrino would modify the shape of the KURIE
plot at low β-particle energies. If the electron neutrino mass is zero, linearity ap-
plies. If the electron neutrino has a mass, linearity is lost and the difference be-
tween the line and the experimentally determined shape of the curve points to a
specific mass of the electron neutrino.
F(Z,E ) p2
_______
_ n(p)
_ __
counts
_____
m >0 m e= 0
e
(Q - Emax )
max
E
Fig. 8.13: Kurie plot: Graphical representation in terms of counts ≈ {n(p) / F(Z,Eβ) p2 } = f(Qβ −
Eβmax). At β-particle energies approaching zero, the mass of the electron neutrino would modify the
shape of the Kurie plot differently.
Historically, a huge number of β-emitters has been analyzed by measuring the ranges
d of the emitted β-particles, e.g. in air.10 The distance appeared to be proportional to
the β-particle’s energies, i.e. Emax
β ~ dmax . A systematic correlation between transforma-
tion constant λ and dair was derived (SARGENT 1933). The double-logarithmic correlation
is logλ = A⋅log dair + B, with ASARGENT and BSARGENT representing numerical factors.
With Eβ ~ d, eq. (8.18) is derived with analogue coefficients aSARGENT and bSARGENT. In
this version, it represents a correlation between maximum kinetic energy of the β-
particles emitted and the transformation constant of the nuclide transformation
✶
K1 ! K2. The same is true for the Qβ-value when used instead of Emax β because of
The β-particles emitted from a nuclide trespass into the matter surrounding the nuclide. This
could be gas, liquid or solid matter such as metals, water etc. The electrons interact with the com-
ponents of matter, which is atoms, and more specifically, nuclei and electron shells. Details of this
“interaction of radiation and matter” are discussed in Chapter 1 of Volume II. The β-electrons will
successively lose their kinetic energy until rest energy. The maximum traveling distance dβmax (re-
member there is a continuous spectrum of β-particle energies) is determined by Eβmax. At values of,
for example, 0.1 MeV and 1.0 MeV an electron travels about 0.1 m or 3.8 m in air, 0.1 mm or 4.3 mm
in water, and 0.07 mm or 2.1 mm in aluminum, respectively.
216 8 β-Transformations II: β−-process, β+-process and electron capture
When comparing the many β-transforming nuclides, Q-values correlate with the
half-life. For larger Qβ-values, transformation steps proceed faster. This fits per-
fectly with the β-transformation parabolas shown in Figs. 6.13 and 6.14, for exam-
ple. The further the nuclides are from the vertex of the parabola, the steeper the
hillside of the two arms of the parabola become. While the unit of the x-axis is Z ± 1
and thus is linear, the y-axis representing mean nucleon binding energy is expo-
nential. Figure 8.14 compares the “win” in mean nucleon binding energy ΔĒB with
the corresponding half-life of this transformation for all the unstable nuclides cov-
ered by both Figs. 6.13 and 6.14, i.e. for all β-transformations along the isobars of
mass numbers 95 and 96.
Obviously, there is t½ = f(ΔĒB). Similarly, the Q-values are (in general) inversely
proportional to the half-life or directly proportional to the transformation constant,
eq. (8.19); see also Tab. 8.1. This also applies to Eβ (eq. (8.20)), because the value of
Qβ generally determines the value of Eβ according to eqs. (8.7)–(8.9). The larger
Emax
β (or Qβ), the faster the transformations proceed. Figure 8.15 correlates experi-
mental data for the same nuclides as shown in Figs. 8.14, 6.13, and 6.14. Energies
and velocity of transformation are proportional, but not linearly. Small changes in
energy (Qβ or Emax
β ) have an impressive impact on the half-life of the transformation,
which is indicated by the exponent n in eq. (8.20).
1
Qβ ~ ~λ (8:19)
t1=2
n
λ ~ Emax
β (8:20)
.
0,16
A=95, -
.
0,14
A=95, +/
A=96, -
0.12
0,12 A=96, +/
.
0,10
ĒB (MeV)
.
0,08
.
0,06
.
0,04
.
0,02
0
0,00
-3
1,E-03 -2 -1
1,E-01 0 1
1,E+01 2 3
1,E+03 4 5
1,E+05 6 7
1,E+0
log(t½ in s)
Fig. 8.14: Successive gain in ΔĒB in β-transformation along parabolas for mass numbers A = 95 and
96 correlates with half-lives t½. For values between the gain in successive mean nucleon binding
energy ΔĒB > 0.1 MeV, half-lives are seconds or less, while for ΔĒB < 0.1 MeV, half-lives approach
hours, days and years.
initial to several possible nuclear states, as shown in eq. (8.14). If transition rates
for all individual states n are all taken together, integrals cover a range of p = 0 to
pmax and result in eq. (8.21), shown in Fig. 8.16.
The right part of eq. (8.21) creates an integral representing the FERMI correction
function F = f(Z,Eβ) and the term of overall transformation and β-particle kinetic en-
ergies f(Qβ –Emax
β ). The latter becomes the “Fermi integral function”, with f as a
function of Z and Qβ. It reflects the impact the individual nuclear structure of the
nucleus has on the transformation process. The constant C and the matrix transition
element, both energy-independent, are not part of the integral, eq. (8.22a). Integrat-
ing n(p)dp corresponds to the overall probability of transformation processes, i.e.
represents all the individual transition states with their individual energetic levels
and individual transition probabilities, eq. (8.22b). It gives λ = f C {Mfi}2. When using
half-life instead of transformation constant, it is ln2 / t½ = f C {Mfi}2, eqs. (8.23a,b).
Parameters of energy (Qβ and Emax β ) are correlated with internal structures of the
218 8 β-Transformations II: β−-process, β+-process and electron capture
1 ms
2 A=95, -
log t½
A=95, +/
A 96, -
A=96,
0 A=96, +/
1s
g( in s-1)
-2 1 min
log
1h
-4
1d
-6
1a
-8
-0.1 0.1 .
0.3 0.5 0.7 0.9 1.1 1.3
log(Emax in MeV)
Fig. 8.15: Correlation of maximum kinetic energy of the emitted β-particles and the transformation
constant (a) and half-life (b) of the β-transformation for nuclides of mass numbers A = 95 and 96.
For changes in Emax
β of one order of magnitude (e.g. from 1 to 10 MeV), the transformation constant
changes by seven orders of magnitude.
nuclides described by quantum physics and now bridge with transformation veloci-
ties λ or t½.11
In addition, eq. (8.24) shows the product of f and t½. It is usually written as
“ft” and gives ft = ln2 / C {Mfi}2. Here, the value of the FERMI integral function f is
multiplied by the half-life t½ (in seconds). Typically, it is expressed as logarithm
log(ft). For each β-transformation of nuclides ✶K1 ! K2, these values are tabulated.
The data for selected unstable nuclides undergoing β-transformation are listed in
Tab. 8.6.
This becomes the rationale of the semi-empiric correlations of type log(λ) vs. log(Eβmax) within
the “historical” SARGENT plot.
8.5 Selection rules 219
pmax pmax
(8.21)
n(p)dp = C Mfi 2 F(Z E )p2 (Q - E )2dp
F(Z,E
p=0 p=0
C and energy-independent M not part of the integral
pmax pmax
(8.22a)
(8.22b)
p=0 p=0
3a)
(8.23
= f C Mfi 2
= ln2 / t½
(8.23b)
ln2 / t½ = f C Mfi 2
f t ≈ f t½
(8.24)
ft = ln2 / C Mfi 2
Fig. 8.16: ft-values: The Fermi integral function f linked to the half-life of β-transition processes.
The ft-values thus combine quantum physical characteristics of transitions and the half-lives of the
transition and are tabulated for each β-transformation.
Until this moment, probabilities of transitions have been discussed for β-transformations
with respect to energetic levels of initial and final states. However, something more is
needed to quantitatively understand β-transformation processes. In quantum phys-
ics, but also in physics and chemistry in general, there are so-called rules of transi-
tion, also called selection rules.12 The fundamental point of view here is “symmetry”.
Symmetry belongs to the fundamental laws of quantum mechanics.
8.5.1 Spin
Rules like these also work in chemistry and are relevant for understanding, for example, elec-
tronic spectra (i.e. different selection rules apply to transitions of electrons between s-s, d-d and f-f
atomic orbitals compared to p-s etc. transitions), or to make use of infrared and RAMAN spectros-
copy (due to selection rules in vibrational states).
220 8 β-Transformations II: β−-process, β+-process and electron capture
all objects defined by orbital momenta – including the photon.13 Changes in overall
spin of a certain nuclear state of a nucleus between the initial and the final state may
according to spin-orbit coupling values J cover all values between 〈Ji –Jf〉 and Ji + Jf.
li − lf ≤ l ≤ li + lf (8:25)
8.5.2 Parity
N S
-x x
-yy y
-z z
mirror
S N
Fig. 8.17: Spatial inversion of a quantum physical object may occur as in a mirror (N = north,
S = south). Parity symmetry would create a further inversion.
. . . which will become an important issue when discussing secondary transition processes in
Chapter 11.
An analogy to chemistry may be found in chirality, when a given molecular structure appears as seen
in a mirror and left-hand arrangements turn into right-hand ones. It remains the same molecule, but its
constituents are arranged conformationally different in space – and create different chemical properties.
8.5 Selection rules 221
involved in wave functions, eigenvalues of the wave functions change, which is ex-
pressed by parity Π. The polar coordinates x, y and z are involved in wave functions
Ψ. Mathematically, parity refers to how wave functions with corresponding eigen-
values and parity operators P change in the course of spatial inversion, as shown in
Tab. 8.3. The process is thus either invariant (i.e. symmetric) or not. If not, it is
called parity inversion (or transformation) or parity violation.15 If the system is in-
variant to inversion and changes into itself, the parity is +. In the opposite case, the
system is variant to inversion and changes, symmetry is violated and the parity is −.
Πi
= ð − 1Þn (8:26)
Πf
Tab. 8.3: Parity: Changes in wave functions of initial to final states follow eigenvalues
corresponding to the parity operator P, parity Π, with different impact on symmetry.
For nuclear transitions, symmetry in orbital and quantum spin and in parity influ-
ence the processes of nuclear transformation. Selection rules for transitions be-
tween initial and final states (beyond energy) are expressed mathematically by eqs.
(8.25) and (8.26) for overall spin and for parity. Changes are ΔJ = 0, 1, 2, 3, 4, . . .,
and either ΔΠ = + or −. Overall parameters combine both changes in terms of overall
spin of the nuclides IfiΠ . Note that spin states are defined according to the shell
model of the harmonic oscillator involving spin-orbit coupling.
The impact of symmetry may qualitatively (maybe oversimplified) be compared
to effects well established for chemical reactions (see Tab. 8.4). Of course, the most
important parameter is energy, for example, the degree of exothermic character. The
mechanism of the chemical reaction, however, may be influenced be some hindran-
ces: steric, kinetic, etc. The lower the hindrance, the more favorably the reaction
While quantum field theory implies that parity violation contradicts fundamental laws in phys-
ics and cannot appear, this is true for strong force, electromagnetism and gravity – but not neces-
sarily true for the weak interaction. In 1957 S E Wu et al. studied effects of 60Co beta transformation
and experimentally revealed violation of parity conservation.
222 8 β-Transformations II: β−-process, β+-process and electron capture
proceeds.16 For nuclear transitions, this is paralleled by the Q-value of the transfor-
mation. In addition, selection rules apply: whenever a change in overall spin or par-
ity is zero or high (ΔJ and/or ΔΠ), β-transition occurs straightforwardly or is hindered.
These transitions are either “allowed” or “forbidden”. One may say that the more
changes involved in overall spin and parity, the more difficult (more forbidden) the
nuclear transformation is.
Tab. 8.4: Qualitative comparison of selection rules in nuclear sciences and in chemistry.
Example: The β−-transformation of tritium 3H, i.e. 1H2 ! 2He1, is of JΠi = s½ from ini-
tial state to JΠi = = s½ for final state (see Fig. 8.18). Parities are + for both the initial
and final states. Consequently, transitions are of ΔJ = 0 and ΔΠ = +.17
In the context of symmetry, changes in overall spin and parity decide whether the
transition process is straightforward – or inhibited. The termini used are “allowed”
and “forbidden” with internal gradations.
– “Allowed”: Allowed transitions are either “super-allowed” or just “allowed”.
Super-allowed refers to the absence of changes in overall spin and parity, i.e.
ΔJ = 0, and ΔΠ = +. They overlap with “allowed” transitions, which still remain
ΔΠ = +, but may accept the lowest change in overall spin: ΔJ = 1.
– “Forbidden”: The more changes there are in ΔJ, the more the transitions be-
come forbidden. The lowest graduation is the retention of ΔΠ = +, but an in-
creased change in spin: ΔJ = 1 or 2. This is “first forbidden”. Further progress in
For example: “similia similibus solvuntur”, similar substances will dissolve similar substances.
The same set of parameters is true for the transition of a free neutron to a proton. The more
nucleons involved, the more complex the determinations of overall spin values of individual nu-
clear states. For 3H the unpaired proton at 1s level contributes a spin of s½, while the two paired
neutrons at 1s level add spins of +½ and −½ to result in an overall spin of ½. For the product nu-
cleus 3He, the two paired protons combine with one unpaired neutron. The remaining overall spin
is again ½.
8.5 Selection rules 223
1d 3/2 4
3
1H2
2s 1/2 2
1d 5/2 6
1p 1/2 2
1p 3/2 4 S=½
(0neutron
1s 1/2 2 + ½proton)
1d 3/2 4
3
2He1
2s 1/2 2
1d 5/2 6
1p 1/2 2
1p 3/2 4 S=½
(½neutron
1s 1/2 2 + 0proton)
Fig. 8.18: Overall spin and parity values for the β−-transformation of tritium.
ΔJ ΔΠ
Allowed , + (no)
Super-allowed (n = )
n s n ! p s/+ ! s/+ No .
H . a H ! He s/+ ! s/+ No .
C min C ! B p/- ! P/- No .
O . s O ! N + ! + No .
S . h S ! P s/+ ! s/+ No .
Allowed (n = )
He . s He ! Li + ! + No .
C . s C ! B + ! + No .
F min F ! O + ! + No .
P . days P ! S s/+ ! d/+ No .
S . days S ! Cl d/+ ! d/+ No .
Fe . min Fe ! Mn f/- ! f/- No .
First forbidden (n = )
Kr . a Kr ! Rb g/+ ! f/+ Yes .
Cd ⋅ a Cd ! In h/- ! g/+ Yes .
I . h I ! Xe g/+ ! h/- Yes .
Second forbidden (n = )
Na . a Na ! Ne + ! + No .
Zr .a Zr ! Nb d/+ ! g/+ No >.
Tc . a Tc ! Ru g/+ ! d/+ No .
Cs . a Cs ! Ba g/+ ! d/+ No .
Fourth forbidden (n = )
K . a K ! Ar ! + Yes .
8.5 Selection rules 225
100 0 + 1+ 100
ΔJ = 0 or 1
ΔJ = 0 or 1; ΔΠ = no ΔΠ = yes ΔJ = 2
super- ΔΠ = yes
allowed
50 50 second
0+ 0+
third
fourth
0 0
0 5 10 15 0 5 10 15
Fig. 8.19: logft-values for unstable nuclides undergoing β-transformation grouped according to
allowed and forbidden transitions.
Thus, it is not only the half-life (in seconds), it is also the impact of symmetry
according to quantum mechanics which characterizes the transformation. For exam-
ple, the half-life of sodium-22 of 2,603 years is shorter than the half-life of tritium of
12,323 years, but ft values are much larger for 22Na (logft = 15.1) than for 3H (logft =
3.1). The transition of 3H as illustrated in Fig. 8.18 is “super”-allowed, the one of 22Na
is “second” forbidden. The changes in ΔJ are responsible and these are 0 and 3, re-
spectively (see Tab. 8.5). Note that there is no change in parity for both transitions.
Table 8.6 summarizes some unstable nuclides undergoing β-transformation.
Historically, some β−-emitters available within the natural decay chains have been analyzed in
that way.
226 8 β-Transformations II: β−-process, β+-process and electron capture
-1
-2
2
207Tl 208Tl
214Pb
-3 211Pb 214Bi
228Ac
-4
212Bi
234Pa
-log( in s-1)
-5
210Bi
-6 234Th
-7
-8
210Pb
-9
-10
-1.5 -1 -0.5 0 0.5 1
log(Emax in MeV)
Fig. 8.20: β-emitting nuclides originally utilized to create the Sargent plot. The two lines
correspond to allowed and forbidden (dotted line) transformations.
Quantum physics provides an explanation. The impact of wave functions and tran-
sitions expressed via {Mfi}2 considers violation of symmetry. Whenever transitions
are forbidden they proceed much slower than an allowed transition – despite the
same balances in energy in Qβ-values or maximum kinetic β-particle energy. The
unstable nuclides illustrated in Fig. 8.20 belong either to those with allowed transi-
tions or to those with forbidden ones.
8.6 Excited states 227
So what about the ground state level of 60Ni? It is not populated directly from 60Co due to selec-
tion rules. This ground state is finally achieved by another mechanism: de-excitation along excited
nucleon shell levels. The process is similar to those introduced for energetically excited electrons
existing at higher-energetic electron shell orbitals. A further similarity is that the most common ap-
proach here is emission of electromagnetic radiation, i.e. photons. This is the topic of Chapter 11 on
secondary transformation pathways.
228 8 β-Transformations II: β−-process, β+-process and electron capture
60Co
*K1 5+
eV
eV
5 271 a
5.271
eV
0.12% / 1491 ke
0.002% / 664 ke
99.88% / 317 ke
100%
Q
2.824 MeV
<0
7.5
3K2 4+
>13.3
2K2
2+
?
14.7
1K2
2+
60Ni
stable gK2 0+
Fig. 8.21: Excited levels following the β−-transformation of 60Co. Primary transition processes of
60
Co populate three separate excited levels ⊙iK2 in the product nucleus 60Ni, indicated by gray
color. The population of the levels follows symmetry considerations. The primary transformation
populating the first excited state is the most abundant one because the changes in spin ΔJ = 1 are
the lowest relative to all the other options. The direct primary transformation into the ground state
is strongly forbidden and does not occur at all (ΔJ = 5). De-excitation from one excited state to an
energetically lower one (“?”) requires secondary transformation processes.
Tab. 8.7: Excited levels in the β−-transformation of 131I (t½ = 8.0233 d, 7/2+,
Qβ = 0.971 MeV) to 131Xe (stable).20
Level Emax
β Abundance ΔJ ΔΠ Category logft
(keV) ( %)
The type β-transformations comprise most of the unstable nuclides. Among these
nuclides are many that have become valuable for research and technology, for in-
dustrial and medical applications. Table 8.8 gives a selection of relevant nuclides.
All the three subtypes of transformations provide different kinds of emitted radia-
tion that is of benefit for specific applications. In some cases, the β−-particle is the
main tool. In many cases, the secondary effect is more important for practical applica-
tions than the primary transformation. The secondary effect of interest is the emission
of low- or high-energetic photons resulting from de-excitation of excited nucleon lev-
els populated within the primary transformation process, discussed in Chapter 11. In
the β+-process, the positrons emitted usually serve as a source of photons, which are
created in the course of annihilation of that positron after combining with an electron.
This effect belongs to the class of post-transformation processes to be discussed in
Chapter 12.
All these applications will also be discussed separately in Chapter 11 according to
the various directions of utilization of radionuclides and their emissions in research
and technology. Here, the most relevant nuclides undergoing β-transformation are
only mentioned with respect to their β-transformation details.
In the case of low-energy and relatively long-lived nuclides, β−-particles are of value
for several purposes. Well-known examples are tritium and 14C. Tritium or 14C-labeled
analogue molecules are used to investigate chemical reaction mechanisms and to an-
alyze biological processes of relevant organic molecules (hydrocarbons) in vitro.
Tab. 8.8: Selected unstable nuclides undergoing β-transformation relevant for applications in
research and technology.
β−-subtype
H . a . – Chemistry, in vitro molecular biology
and assays, dating
C a . –
P . d . – In vitro molecular biology and assays
S . d . –
Co . a ., . . External radiation therapy
Sr . d . – Nuclear medicine therapy
Y . . –
Mo h . – Parent of mTc generator
Cs . a ., . . Calibration source
Sm . d ., . . Nuclear medicine therapy
Lu . d . .
Re . h . .
Re . h . .
β -subtype
+
C . min . – Nuclear medicine diagnosis
N . min . –
O . min . –
F . min . –
Al .⋅ a . ., . Astrophysics
Ga . min . – Nuclear medicine diagnosis
I . d . ., .
ε-subtype
Co . d – ., . MÖSSBAUER spectroscopy
I . h – . Nuclear medicine diagnosis via SPECT
I . d – . In vitro molecular biology and assays
Er . h – – Nuclear medicine therapy
8.7 Examples and applications 231
These nuclides do exclusively emit beta particles and can be detected by liquid scin-
tillation. This deserves separate discussion; see Chapter 12 of Volume II.
In tumor therapy, β−-particles are employed to induce dense ionization of water
within or around a tumor cell. Here, a key nuclide is 90Y. It transforms almost
completely (99.983 %) into the ground state of 90Zr according to the logft-value of
8.0 of the 2− to 0+ transition, shown in Fig. 8.22. Consequently, the β-particles emit-
ted carry almost the complete amount of energy of the transition (Qβ − = 2.280 MeV)
and the Emaxβ value is identical, namely 2.280 MeV. Since there are two almost un-
populated excited levels (⊙2K2 with logft = 11.1 and 1.4⋅10−6 %, ⊙1K2 with logft =
9.4 and 0.017 %), there is hardly any accompanying photon emission due to de-
excitation. Thus, 90Y is an almost “pure” β−-emitter of high-energy β−-particles of
long range in tissue (maximum range = 12 mm, mean range = 5 mm) and is preferred
for treatment of tumors of larger dimension.
90Y
*K1 2-
keV
0.017% / 519 keV
64.10
64. 0h
eV
11.4 10 % / 94 ke
99.983% / 2280 k
Q
2.280 MeV
11.1
2K2 2+
9
9.4
1K2 0+
8.1
90Zr
gK2 0+
stable
Fig. 8.22: β−-transformation of 90Y to 90Zr. Direct transformation to the ground state is the
dominant strategy (99.983 %) due to the logft-value of 8.0. There is negligible population of two
excited states.
Once a biological carrier system has delivered the radionuclide close to the tumor
cell or even inside the cell close to the cell nucleus, ionization caused by the emitted
232 8 β-Transformations II: β−-process, β+-process and electron capture
β−-particles21 leads to double-strand breaks of the tumor DNA and selective death of
the malignant cell. This approach may be tuned to large or small tumors according to
the kinetic energy of the β−-particles, which is proportional to their range in tissue, as
shown in Fig. 8.23.
Emax
2.3 90Y E max (mean)
( )
= 2.28 (0.9) MeV
90Y
2.2 particle range in water = 12 (5) mm
188Re
2.1
2.67 d 2.0
3.19 h 2.3 1.9
(2186) 1.8
1.7
17
1.6 186Re
1.5
131I 1.4 166Ho
8.02 d 1.3
0.6, 0.8:, ... 1.2
364,, 637,, … 11
1.1 153Sm
1.0
0.9 131I
177Lu
0.8
0.7
0.6 177Lu
160.1 d 6.7.1 d 0.5
0.5, 161Tb
0.4
208
0.3
0.2 169Er
Fig. 8.23: Medically relevant β−-emitters providing β−-particles of short to long range in water.
β−-emitters such as 90Y, 131I and 177Lu provide β−-particles of varying maximum kinetic energies.
These β−-particles induce dense ionization of water within or around a tumor cell, which ultimately
leads to double-strand breaks of the tumor DNA and selective death of the malignant cell. By
selecting the appropriate β−-emitter, the range in tissue may be controlled. The lanthanide 177Lu is
among the very useful trivalent metallic radionuclides because of the low mean range of its
β−-particles in tissue of 0.67 mm. The dimension of one tumor cell is 10–20 μm.22
The effects induced by the various kinds of emissions when interacting with matter are dis-
cussed in Chapter 1 of Volume II, and nuclear medicine therapy is discussed in Chapter 14 of Vol-
ume II.
Remember that ranges of β−-electrons are not discrete due to the continuous energy spectrum
of the electrons. Mean ranges thus correspond to the energy of most of the electrons initially
emitted.
8.7 Examples and applications 233
β+-electron emitters (i.e. proton-rich unstable nuclides of low mass number A) are of
interest in most cases not because of the radiation caused by the primary nuclear
transformation (which is the positron), but because of post-effects the positron under-
goes once released from the initial nucleus. This is the subsequent annihilation of the
positron when combined with a normal electron. This process transforms the mass of
the two elementary particles into electromagnetic emission in the form of two 511 keV
photon quanta, see Chapter 12. These photons are detected by Positron Emission To-
mography (PET). The nuclides providing intense and almost exclusive positron emis-
sion, such as and 18F, are labeled to biologically relevant molecules and thus are
234 8 β-Transformations II: β−-process, β+-process and electron capture
99Mo
*K1 1/2+
keV
keV
keV
0.134% / 353 keV
0.111% / 215 eV
0.0016% / 185 eV
82.1% / 1215eV
0.052% / 686 keV
keV
0.0021% / 158 k
1.18% / 848 k
16.45% / 437 k
k
k
0.0027% / 285 k
k
100%
Q
2.748 MeV
3/2+
8.7 8.9
3/2+
7.4
3/2+
8.55
8
1/2-, 3/2-
9.4
5/2+, 7/2+
iK2 8.0
3/2-
6.2
1/2+
9.5
3/2-
11.0
3/2+
8.4
99mTc 3/2-
7.1
6.007 h mK2 1/2-
99gTc
gK2 9/2+
2.11.105 a
Fig. 8.24: β−-transformation of 99Mo to 99gTc and 99mTc. Direct transformation to the ground state
is forbidden. There are many individual excited levels available, but one preferred pathway (82.1 %)
with logft of 7.1. Unlike the processes shown for 131I, this excited state exhibits its own half-life of
6.0 h and thus represents the metastable nuclide 99mTc.
valuable tools for molecular imaging in nuclear medicine.23 Despite their promising
nuclear data, they are particularly relevant in radiopharmaceutical chemistry and
molecular imaging, because these nuclides represent isotopes of “organic” atoms,
i.e. those typically constituting organic and biologically essential molecules.
8.8 Outlook
8.8.1 De-excitation
In many cases, nucleon transformation does not cover the complete process of
✶
K1 ! gK2; but rather the route ✶K1 ! ⊙K2. This de-excitation procedure requires
transitions:
(a) between excited states and
(b) from exited states to the ground state.
Similar to the excited electrons populating electron shells of higher energy, there
must by a process of de-excitation towards the ground state. The most common ap-
proach for the excited nucleus of an atom is emission of electromagnetic radiation,
i.e. photons. Interestingly, this applies to most nuclides ⊙K2 providing excited states,
this means that it does not matter whether the nuclide K2 itself is stable or not (see
the example given in Fig. 8.23 with stable 60Ni). This belongs to the class of “second-
ary” nuclear transformation in terms of ✶K2 ! K2 and is the topic of Chapter 11.
The β-process itself shows some special features: inverse β-transition and double
beta transformation. Nuclear sciences make use of the inverse process of proton-to-
neutron conversion. Instead of the emission of an electron neutrino in the course of
the process p ! n + β+ + νe, the proton may capture an electron antineutrino. This
also creates a neutron (similar to the process of electron capture), but exclusively
releases a positron, summarized in eq. (8.27). The product can be easily and sensi-
tively detected, and thus serves as a measure of whether the electron neutrino was
there or not. This was briefly mentioned in the context of the (indirect) experimen-
tal detection of the electron neutrino itself.
νe + p ! n + β + (8:27)
Despite the standard procedure of converting an excess neutron into a proton (or a
free neutron into a proton), there are two situations in which two “bound” neutrons
simultaneously transform into two protons. The mechanism is either according to a
real “double” β-process, i.e. emitting two β−-particles and two electron antineutri-
nos (see eq. (8.28) and Fig. 8.25), or a “neutrino-less” process (eq. (8.29)). While
the second version is a matter of intense research, the first has been experimentally
236 8 β-Transformations II: β−-process, β+-process and electron capture
observed for about 35 unstable neutron-rich nuclides, such as 48Ca, 76Ge, 82Se, 96Zr,
100
Mo, 116Cd, 128Te, 130Te, 136Xe, and 150Nd.
A✶ −
2νe β β Z K1 ! Z + 2 K2 + 2β
A
+ 2νe (8:28)
A✶ −
0νe β β Z K1 ! Z + 2 K2 + 2β
A
(8:29)
The rationale is, that whenever the “normal”, single β−-transformation process is
not energetically possible or symmetry-wise forbidden, the double process becomes
an option. Its probability, however, is small. The nuclides mentioned thus may
show very large half-lives approaching 1020 years.
PARTICLE
ELECTRON
ANTI-NEUTRINO
PARTICLE
ELECTRON
ANTI-NEUTRINO
Fig. 8.25: Mechanism of a double β−-transformation process. Kinetic energy in this case is
distributed among all four emission products.
Two examples are shown in Fig. 8.26 for the double β-transformation process in 82Se
and 96Zr. The latter nuclide was illustrated within the β-parabola of the A = 96 isobar
chain in Fig. 6.14. Its expected transformation product was 40Zr56 ! 41Nb55. Neverthe-
less, 96Zr is not the nuclide at the vertex of the parabola, so there should be a path-
way to transform to the final and stable nuclide, which is 96Mo (ĒB = 8653.988 keV).
However, mean nucleon binding energies of the two nuclides are ĒB = 8635.328 and
ĒB = 8628.997 keV, respectively, for 96Zr and 96Nb so there is (energetically) no good
reason to transform. The alternative is indeed the double β−-route, although the half-
life of this transformation is 3.9⋅1019 years.
8.8 Outlook 237
96Mo
M 16.68
96Nb
9
23.4 h
96Zr
82Kr 8.73
99.941 _ 3.9.1019 a
2 e 2
96Y
82Br
9.6 s 5.4 s
6.1 m 35.3 h
82Se
8.73
_ 1.08.1020 a
2 e 2
82As
13.6 s 19.1 s
Fig. 8.26: Examples of double β−-transformation processes. “Normal” β−-transformation from 82Se
to 82Br and from 96Zr to 96Nb is “problematic” due to balances in mean nucleon binding energy and
masses of the nuclides involved. Atomic masses (in u) for the nuclides involved are 81.9166995,
81.9168018, 81.9134812 for 82Se, 82Br, 82Kr and 95.9082776, 95.9081016, 95.9046748 for 96Zr,
96
Nb, 96Mo, respectively.
9 α-Emission
Aim: For mass numbers 1 to 209, the β-process transforms unstable nuclides into sta-
ble final product nuclides, forming the valley of β-stability. Whenever heavier nuclides
(Z > 83, A > 210) behave according to the β-transformation profile, the most stable nu-
clide of the A = constant parabola does not necessarily represent a stable product.
Consequently, another option for stabilizing unstable nucleon compositions is
needed, and the second most “popular” one (after β-transitions) is α-emission.
The essence of α-emission is the emission of a cluster of four nucleons, two pro-
tons and two neutrons – the nucleus of the 4He atom, the α-particle. The new nu-
clide K2 thus will be of (A − 4), (Z – 2) and (N − 2). According to the liquid drop
model, mean nucleon binding energy improves for heavy nuclides in the direction
of a lower mass (A − 4) position until medium mass numbers are reached.
This chapter discusses:
(i) how the overall energy is disseminated among the new nuclide K2 and the α-
particle, in particular in terms of kinetic energies of the α-particles emitted,
(ii) the correlation between the nuclear transformation energies and the velocity of
the transformation, i.e. its half-life, and
(iii) the corresponding theoretical and mathematical models. The main feature is
the quantum mechanics tunneling effect.
Finally, prominent α-emitters are introduced, which are important to specific as-
pects of science, technology and medicine.
9.1 Introduction
For all mass numbers from 1 to 209, β-processes yield one definite stable nuclide or
two – depending on the (even, even) or (even, odd) nucleon composition. This para-
digm does not continue after A > 209. For example, let us consider the mechanism of a
β-transformation of the A = 226 isobar. The radium isotope 226 represents the most sta-
ble nucleon composition. 226Ra shows by far the longest half-life of this isobar, namely
1600 years. The neighbors at (Z + i) are of much lower stability, with half-lives the
range of hours (29 hours for 226Ac), minutes (31 min and 1.8 min for 226Th and 226Pa)
and decrease further down to milliseconds (280 ms for 226U and 31 ms for 226Np). For
the (Z – i) arm of the parabola, 226Fr and 226Rn show half-lives of 48 s and 7.4 min,
respectively. In the present case, the nuclide at the vertex of the isobar parabola
of ĒB = f(Z) is 226Ra, yet it is not stable, as shown in Fig. 9.1. Thus, β-transformation
has done its best to build the most stable nuclide of the A = 226 isobar, but was not
able to create a stable nucleon configuration. Consequently, 226Ra must transform
into a more stable nucleon configuration by a different mode to the β-transformation.
https://doi.org/10.1515/9783110742725-009
240 9 α-Emission
226
6Np
31 ms
226U
0 28 s
0.28
226Pa
1.8 min
226Th
31 min
226Ac
29 h
226Ra
1600 a
226Fr
48 s
222Rn 226Rn
Fig. 9.1: β-transformation processes along the isobar A = 226. The most stable, but not really
stable, only just-stable and longest-lived nuclide is 226Ra. This unstable nuclide of optimum mean
nucleon binding energy along the isobar transforms through α-emission to 222Rn, thereby
switching to a new, lower isobar. In the Karlsruhe Chart of Nuclides, α-emitting radionuclides are
indicated by yellow color.
9.2 Mass balances in α-transformations 241
Emission of an α-particle immediately reduces the mass of the unstable nuclide ✶K1
and changes both proton and neutron numbers – it is a primary transformation.
The reason the small component is the α-particle and not any other mixture of nu-
cleons lies in its very high “internal” stability. The mean nucleon binding energy of
the 4He nucleus is 7.052 MeV, and the nucleus is further stabilized due to a double
magic nucleon shell configuration (Z = 2, N = 2).
The α-transformation thus balances mass between the initial unstable nuclide
✶
K1 and the transformation product nuclide K2 in a clear way: the mass number of
the new nucleus is lowered to (A − 4). At the same time, the proton number de-
creases by (Z − 2), and the newly formed nuclide K2 belongs to a chemical element
located not as a neighbor, but at (Z – 2) position, as per Fig. 9.2. All α-emissions
follow an isodiaphere line.
✶
K1 ! K2 + α + ΔE (9:1)
A✶
2 K1N ! AZ −−24 K2N − 2 + 42 α 2 + ΔE (9:2)
N
Fig. 9.2: Principal scheme of α-emission within the Chart of Nuclides. The initial unstable nuclide
✶
K1 emits an α-particle, and the new nuclide K2 is of (Z − 2), (N − 2) composition. α-emission
follows an isodiaphere line.
242 9 α-Emission
The “motivation” to transform via α-emission is the same as for any other pro-
cess of radioactive transformation, namely to increase mean nucleon binding en-
ergy. This is directly achieved for large values of A when shifting 4 mass units to
the left, as per Fig. 9.3. Within the LDM, the main impact originates from the de-
crease of the value of the coulomb term of the WEIZSÄCKER equation, which is –γLDM
Z2 / A⅓. In contrast, the asymmetry term distributes less than the value of (N – Z),
for ✶K1 is identical with that of (N − 2) − (Z − 2) for K2 after α-emission. Also, the sign
of the parity terms does not change within the course of the α-emission: any (even,
even)-nucleus will remain (even, even) after transformation, any (even, odd) or
(odd, even) composition will remain the same, etc.
.
9,0
.
8,5
A-4
.
8,0
ĒB (MeV)
.
7,5
.
7,0
LDM A LDM A⅔ LDM
Z2 LDM
(N-Z)2 LDM
EB = - - - ±
A⅓ A A¾
.
6,5
.
6,0
0 50 100 150 200 25
A
Fig. 9.3: Directions of α-emission processes in terms of improving mean nucleon binding energy
according to the LDM. The line gives the polynomial of the Weizsäcker equation, and the squares
represent the mean nucleon binding energy of stable nuclides.
9.3 Pathways of α-emission 243
Figure 9.1 introduced the scenario of nuclide stabilization at mass numbers A > 209
along an isobar line, i.e. for β-transformation pathways. In these cases, the local
maximum in ĒB for A = constant does not represent a stable nucleon configuration,
and consequently, another transformation pathway is needed.
However, the emission of one α-particle again does not necessarily generate a stable
nucleon mixture. This effect can be explained following the example given in Fig. 9.1.
226
Ra is a very neutron-rich nuclide. The excess of neutrons is 138 − 88 = 50, the ratio
between neutrons and protons is 138:88 = 1.568. Whenever 226Ra starts to transform
by α-emission it must follow an isodiaphere line, forming a product of (Z – 2) and
(N – 2) composition. Thereby, the absolute excess of neutrons remains constant, but
the relative excess of neutrons and the ratio of neutron-to-proton number further in-
creases. The transformation product is 222Rn with 86 protons and 136 neutrons; the
neutron excess remains the same (50), while the ratio increases (136:86 = 1.5814).
Consequently, the new nuclide is not expected to be stable. Again, this single
α-emission transformation is not able to produce a stable transformation product, al-
though there definitely is a reduction in mass: the mass of 222Rn is 222.017578 u and
is thus less than the mass of 226Ra of 226.025410 u by the mass of the α-particle. This
is accompanied by a gain in mean nucleon binding energy of ĒB = 7.662 MeV for
2226
Ra and 7.695 MeV for 222Rn.
So what? The transformation may continue via the next α-emission. This is ex-
actly the case for 226Ra, as introduced for the natural chain of transformation of the
4n + 2 series. 226Ra originates from 230Th by α-emission, and 226Ra itself continues to
form daughters by successive α-emission as 226Ra ! α ! 222Rn ! α ! 218Po ! α !
214
Pb, shown in Fig. 5.23. With this nuclide, the α-chain finally terminates; the ex-
cess of neutrons has reached a high level.
Now, here comes the “teamwork” of α- and β − -transformations: for 214 Pb, the
β−-process becomes the only pathway to further stabilize the nucleus. For this natu-
rally occurring decay chain, this β−-transformation now represents the continuation
of the preceding α-emission step(s). It happens along the neutron-rich arm of the iso-
bar parabola at A = 214, until a new, local maximum in ĒB is reached for this particular
244 9 α-Emission
isobar. This new maximum could represent a stable nuclide; if not, a situation occurs
like that explained above for transformations along the isobar A = 226, and another
α-emission follows, as per Fig. 9.4.
226Ra
1600 a
214Rn 222Rn
0.3
<0 3 s 3 825 d
3.825
214At
< 0.8 s
214Po 218Po
214Bi
19.9 min
210Pb 214Pb
210Tl
1.30 min
Fig. 9.4: Continuation of the naturally occurring 238U transformation chain subsequent to the
α-emissions from 226Ra. The direct chain of α-emission stops at 214Pb to continue along the isobar
214 via β−-processes.
Nature has realized this within the naturally occurring chains of transformation.
9.3 Pathways of α-emission 245
Fig. 9.6: Notation of parallel options of primary transformations for one and the same nuclide.
The size of the color-coded area qualitatively indicates the proportions between the different
branches of the transformation. 212Bi: 35.93% α-emission (Qα = 6.207 MeV, main α-energy
6.167 MeV) + 64.07% β−-emission (Qβ = 2.252 MeV, main β−-energy 2.252 MeV). 211At: 41.78%
α-emission (Qα = 5.982 MeV, main α-energy 5.87 MeV) + 58.2% electron capture (ε) emission
(Qε = 785 keV).
nuclide!2 This indicates that Q values are positive for different primary transformation
options. In this case, each pathway gets its individual absolute value according to the
different balances in mass, and the notations are Qα and Qβ, respectively. The branch-
ing of the two transformations is specific. Figure 9.6 shows examples of a nuclide gen-
erated in a naturally occurring chain of transformations utilizing β−-transformation in
parallel to α-emission (212Bi) and an example of an artificially produced radionuclide
performing both α- and electron capture transformations (211At). Prominent examples
are listed in Tab. 9.1.
This simultaneous transformation is the reason for the branched chains within the four naturally
occurring chains of transformation.
246 9 α-Emission
210Po 214Po
138.38 d 164 ms
210Bi 214Bi
5.013
5 013 d 19 9 min
19.9
206Tl 210Tl
206Hg
8
8.15
15 min
Fig. 9.7: The terminal stable nuclide 206Pb within the naturally occurring 238U chain. Either the
α-transformation forms a stable nuclide, i.e. 210Po via α ! 206Pb, or the β−-transformation of 206Tl
via β− ! 206Pb does so.
9.4 Energetics
The α-transformation process occurs spontaneously and is nonreversible like all the
other primary transformation pathways. Thus, α-transformation is exothermic, and
the loss in mass (lower mass for the sum of the transformation products (mK2 + mα)
relative to m✶K1) corresponds to energy ΔE, as per eq. (9.3). The corresponding energy
may also be derived using the mass defect Δmdefect or Δmexcess values of the three
components involved, as per eq. (9.4). The value of ΔE may specifically equal the
Qα-value of the process.
9.4 Energetics 247
9.4.1 Values of Qα
Figure 9.10 illustrates the regions these α-emitters populate within the Chart of Nu-
clides, and Tab. 9.1 list several important α-emitters.
248 9 α-Emission
1 2 3
8
Z = 82
82
26
8
N=8
N = 12
4
Q (MeV)
Z = 82
2
-2
-4
100 120 140 160 180 200 220 240
A
Fig. 9.8: Values of Qα for unstable nuclides not undergoing β-transformation. There is a general
increase in Qα with increasing mass number. At the same time, several mass numbers show a
significantly increased Qα-value. This reflects the impact of magic numbers of protons and
neutrons. (a) N = 82, (b) Z = 82, (c) Z = 82 or N = 126. The corresponding mass numbers are
A ≈ 144–148 (N = 82), A ≈ 184–192 (Z = 82), and overlapping A ≈ 210–219 and 206–218 for N = 126
and Z = 82.
9.4.2 Values of Qα
A = 144-148 A = 210-219
148Gd 215Ac
A 217Ac
A
214Ra 216Ra
212Rn 214Rn
210Po 212Po
N= 82 84 126 128
A = 184-192 ...
83
A = 206-218
83
Fig. 9.9: Impact of the shell model of the nucleus on Qα-energy. Whenever an α-emission creates a
nuclide K2 of magic nucleon number, the mean nucleon binding energy increases. The
corresponding mass numbers representing large Qα-values are thus around A ≈ 144–148 (N = 82),
A ≈ 184–192 (Z = 82), and A ≈ 210–219 and 206–218 for N = 126 and Z = 82, respectively.
250 9 α-Emission
114
nuclides > ₈₃Bi
50 limited number
of lanthanides
82
50
Impulse p = m⋅v and kinetic energy E = ½m⋅v2 can be found for the two species
formed in the primary nuclear transition, resulting in a balance between α-particle
emission and K2 recoil of mα·Eα = mK2·EK2. The overall energy Qα is allocated to the
emitted α-particle and the recoil nucleus RECOILK2 according to eqs. (9.5)–(9.7).3
Table 9.1 gives the corresponding numbers for the α-emission of 238U. Table 9.3 gives
typical values for selected mass numbers and α-particle energies of 1, 3 and 6 MeV.
pK2 = pα (9:5)
Qα = RECOIL
EK2 + Eα (9:6)
Qα
Eα = (9:7)
1 + mmα
K2
This is different to the β−- and β+-subtypes of the β-process, but similar to electron capture.
9.4 Energetics 251
Tab. 9.1: Important nuclides undergoing α-transformation, their proton number, the transformation
product nuclide, main✶ kinetic energy Eα, half-lives, and (if applicable) parallel primary
transformations (sf = spontaneous fission, see Chapter 10).
✶
Z K K Eα (MeV) t½ Other primary transformation
Gd Sm . . a –
Tb Eu . . h ε, β+
Bi Tl . min β−
Bi Tl . . min β−
Po Pb . . d –
At Bi . . h ε
Ra Rn . . d –
Ra Rn . . d –
Ra Rn . a –
Ac Fr . . d –
Th Ra . a –
Th Ra . .⋅ a sf
U Th . .⋅ a –
U Th . .⋅ a sf
U Th . .⋅ a sf, β−
Np Pa . .⋅ a sf
Pu U . . a sf
Pu U . .⋅ a sf
Am Np . . a sf
Cm Pu . a sf
Bk Am . a –
Cf Cm . . a sf
✶
Main α-emission: Similar to β-transformation, there may be several emissions for the same
nuclide due to the population of excited nuclear states of the product nucleus. The table shows the
most abundant one, which is not necessarily the most energetic one (if primary transformation
prefers an excited state of the product nuclide).
For the naturally occurring transformation chains, for example, most of the α-emissions occur at
A > 208, and the parent nuclides are 232Th, 235Ac and 238U. Consequently, the ratio between masses
of the α-particle and the recoiled nucleus K2 are of 4 to 204–238 and thus relatively similar, i.e. 1
to ≈55 ± 5. This ratio of mK2/mα is completely different compared to the balances of the β-processes:
Masses of nuclides for A = 10 or A = 100 were about six orders of magnitude larger than the rest
mass of the β-particle.
252 9 α-Emission
3. Consequently, kinetic energies of the α-particle are discrete. Figure 9.11 shows
an α-spectrum of the nuclide 148Gd (t½ = 74.6 a) with the discrete energy of the
3.183 MeV α-particle emitted.
4. Compared to the (maximum) kinetic energies of β−- and β+-particles, (discrete)
kinetic energies of α-particles are relatively high. Kinetic energies of α-particles
emitted from the most relevant α-emitters are in the range of 3–6 MeV, although
extreme values of ca. 2 and ca. 12 MeV can be reached.
5. The recoil energy of the product K2 is RECOILEK2 = Qα − Eα, as per eq. (9.6).
6. The recoil energy of K2 is higher if: the kinetic energy of the α-particle emitted
is high and the mass number of the nuclide is low. Recoil energy may typically
range from ca. 10 to 100 keV, as per Tab. 9.3.
Tab. 9.2: Values of Qα and kinetic and recoil energies of the transformation products of the 238U
α-emission process. Mass excess data are used for determination of Qα. Kinetic energies of the
α-particles according to eq. (9.7) are obtained by simply using mass numbers.
Similar to the many β-transforming nuclides, Qα-values correlate with the half-life of
the radionuclide, as per eq. (9.8). The larger the Qα-value, the larger the gain in en-
ergy a nuclide “wins” in terms of mean nucleon binding energy ΔĒB when transform-
ing. In these cases, the transformation steps proceed quickly. The same principle
applies to small Qα-values and/or transformations with special nucleon shell configu-
rations involved. The α-transformation thus shows – similar to β-transformation –
half-lives covering milliseconds to billions of years.
Similar to the SARGENT graph for β-emissions, there is a plot for α-emissions (GEI-
GER / NUTTALL, back in 1911, 1912). Here, dα is the range of an α-particle in air, and
AGEIGER/NUTTALL and B GEIGER/NUTTALL are constants of a linear graph when plotted in
the double-logarithmic version. With a given (exponential) proportionality between
the energy of an α-particle and its range in air (the higher the energy, the greater its
9.5 Velocities of α-transformation 253
200
200
_ 148 Gd
148
150
150 _ FWHM
counts
100 14.8 keV
_
50
counts
100 _ 0
3120 3160 3200 3140
E (keV)
(k V)
_
50 _
0 I I I I I I I I I I
1 2 3 4 5 6
E (MeV)
Fig. 9.11: Experimental α-spectrum of a 148Gd sample over a broad range of energy (0–8 MeV) and
(inset) in the range of 3.10 to 3.25 MeV. The discrete energy of the 3.183 MeV α-particle emitted is
shown. According to the resolution of the α-spectrometer applied and the thickness of the 148Gd
source, this peak has a full width at half maximum (FWHM) of 14.8 keV.5
distance; and dα (in cm) ≈ 0.3 Eα⅔), eq. (9.9) turns into eq. (9.10). Small changes in
energy cause a significant impact on the half-life of the transformation.
1
Qα ð⁓Eα Þ⁓ ⁓λ (9:8)
t1=2
This should not be confused with the initial discrete energy of this emission.
254 9 α-Emission
Tab. 9.3: Kinetic energies of α-particles and K2 recoil energies RECOILEK2 depending
on mass number A of the initial nuclide ✶K1 and Qα-values.
2048
1536
228Th 228Th
count rate (per second)
28.2% 71.1%
1024
0
4500 5000 5500 6000 6500 7000 7500
energy [keV]
Fig. 9.12: Experimental α-spectrum including the four α-emitting members successively formed in
the 232Th chain. The higher the energy, the shorter the half-life. 228Th (t½ = 1.913 a, Eα = 5.340 and
5.423 MeV),6 224Ra (t½ = 3.66 d, Eα = 5.685 MeV), 220Rn (t½ = 55.6 s, Eα = 6.288 MeV) and 216Po
(t½ = 0.15 s, Eα = 6.778 MeV).
Fig. 9.13: Correlation between kinetic energies of the emitted α-particles and the half-life of the
transformation for nuclides present within the naturally occurring transformation chains. The four
chains show more or less comparable values for the constant aGEIGER/NUTTALL in eq. (9.10), but have
different values for the constant bGEIGER/NUTTALL. Again, this hints at the influence of shell effects.
The (4 n + 0) = 232Th and (4n + 2) = 238U chains represent (even, even) nuclei in contrast to the
(4n + 1) = 237Np and (4n + 3) = 235U chains.
The relationship materializes when, for example, correlating the α-particle en-
ergies of four members of the naturally occurring 232Th transformation chain with
their half-lives. Figure 9.12 shows an experimental α-spectrum similar to Fig. 9.11:
228
Th ! 224Ra ! 220Rn ! 216Po. It reflects the relationship between α-particle en-
ergy and half-life: the higher the energy of the α-particle, the shorter the half-life.
This correlation is shown in Fig. 9.13 for selected α-emitting nuclides generated
within the naturally occurring transformation chains.
256 9 α-Emission
e2
UC = Znucleus Zα kC (9:11)
r
Consequently, one must expect that an α-particle leaving this nucleus should have
at least 28.5 MeV energy.8 However, fact 2: the kinetic energy of the emitted α-
particle is 4.198 MeV (only!), as calculated in Tab. 9.2. This is not because it is a
simplified calculation or deduced from a wrong equation – it precisely corresponds
to the experimentally measured kinetic energy of the α-particle as released from
238
U. Both facts are absolutely correct. So, what now?
30
EC COULOMB barrier
20
10
U (MeV)
-10
-20
-30
0 20 40 60 80 100
r (fm)
Fig. 9.14: One-dimensional potential well showing the potential energy U = f(r) of the nucleus of
238
U with the height of the coulomb barrier EC. The energy at the radius of the uranium nucleus of
9.3 fm is ca. 28.5 MeV. With permission: W Loveland, DJ Morrissey, GT Seaborg, Modern Nuclear
Chemistry, Wiley Interscience, 2006, J Wiley & Sons, Inc., Hoboken, New Jersey.
to the strong force at bounded states according to nucleus shell structure) and has
“tunneled” through the potential wall. It becomes a “free” particle. The kinetic energy
of the α-particle after tunneling through the COULOMB barrier is much lower than the
height of the barrier and exactly of the value calculated according to eq. (9.7).
EC
of „COULOMb barrier“
height (extent) E C
E
of -particle emitted
kinetic energy
U(r) 1/r
Nuclear potential U MeV)
quasi-bound states
bound states
ATTRACTION REPULSION
(nucleus potential) (COULOMB potential)
potential well
(strong force)
rNUC
rC
Radius r
EUS
Fig. 9.15: Concept of an α-particle tunneling through a coulomb barrier of a nucleus. EC gives the
amount of the potential energy due to the coulomb forces, Eα the kinetic energy of the α-particle
after tunneling at a virtual radius rC. The smaller rC is, the lower the kinetic energy of the α-particle
emitted.
9.6 Quantum mechanics of α-transformation phenomena 259
A quantum mechanical model (GAMOW 1928, CONDON and GURNEY 1928, 1929) consid-
ers an initial and a final state of transformation, with a corresponding probability
pfi of transition (per unit of time) from one energy eigenstate of a quantum system
(nuclide ✶K1) to another (nuclide K2).11
For α-emissions, the model should allow for three relevant steps:
1. Prior to the emission, the α-particle must be pre-formed as such inside the ho-
mogeneous ensemble of the many individual nucleons within the nucleus. This
may happen with a given probability due to the special stability of a (Z = 2, N = 2)
cluster of double magic shell characteristics, accompanied by a gain in mean
nucleon binding energy.12
2. If the cluster was formed anywhere within the nucleus, this cluster must be
present close to the surface of the nucleus. This includes an anticipated sort of
“transport” or “diffusion”.
3. Only following its formation and virtual transport can the tunneling of the par-
ticle through the barrier be discussed.
Figure 9.16 illustrates the mathematical factors involved in α-particle tunneling. Al-
though there is evidence for the processes of pre-formation of an α-particle (1) and
its diffusion to the surface of the nucleus (2), the mathematical model collects these
two steps into the frequency factor f (also called “reduced transition probability”).
This describes how often an α-cluster appears at the surface of the nucleus (and
“knocks on the door”). Once it has appeared at the surface, it gets a chance to leave
the nucleus via the tunneling effect. The probability for this step is defined by the
penetrability factor P (also known as “transition factor” or “penetrability”).
The two components to be mathematically handled are thus the frequency fac-
tor f and the penetrability factor P. Obviously, both factors determine the overall
efficacy of α-emission. The overall approach is to define the total velocity of an
α-emission as a product of type λ = f⋅P. The individual expressions for the two factors
are summarized in Fig. 9.17.
Despite the fact that the concept is based on quantum mechanics, most of the
parameters involved in eqs. (9.12) and (9.13) reflect basic parameters of the nucleus
of an atom. Parameters used are:
This is quite similar to β-transformations. However, the β-transformation is just based on the
conversion of one sort of nucleon into the other one.
In fact, in situ formation of an α-cluster within a nucleus is known for light nuclides, where it
explains the existence of halo-nuclei. Here, internal clusters are formed with the remaining nucle-
ons occupying “outer shells”. This yields radii for those nuclei that are larger than predicted by
charge or mass radii according to eqs. (2.2a) and (2.2b), e.g. for 6 He and 8 He nuclei with one
α-cluster each and two and four neutrons, respectively, existing at outer spheres.
260 9 α-Emission
1 2 3
Fig. 9.16: The steps involved in the α-emission process. (1) Formation of an α-cluster inside the
nucleus, (2) delivery of this α-cluster towards the surface of the nucleus, (3) release of the
α-particle and tunneling through the coulomb barrier of the nucleus.
The exponent in eq. (9.13) is expressed as 2G, forming eq. (9.14). G is the GAMOW
factor and its values are on the order of G = 30–60. With this dimension, it becomes
clear that the probability of penetrating (tunneling) a potential well is in fact ex-
tremely low. The same is true of the frequency factor. Depending on the proton
number, it is about ZK2−4/3, which makes 1.9⋅10−20 s−1 for ZK2 between 58 and 98.
Primary transformation processes of unstable nuclides ✶K1 do not necessarily yield the
ground state of the newly formed nuclide K2 directly, but may populate energetically
enriched levels of the new nuclide K2. This holds true for the α-transformation as well,
though to a lesser extent than for β-transitions. The number and the characteristics of
9.7 Excited states 261
(9.13)
[2(EC + Q ) / m ]½ 2 2ZK2e2
(9.12)
f= P ≈ exp – 2m´ ( - Q ) dr
2rK2 rK2
RK2
(9.14)
G = GAMOV factor
P = e-2G (2G ca. 60 -120)
(9.15a)
=fP
(9.15b)
t½ = ln2 / f P
[2(EC + Q ) / m´ ]½
(9.16)
t½ = ln2 / e-2G
2rK2
Fig. 9.17: Half-lives of α-emissions correlated with quantum mechanical parameters as well as with
Qα– values in eq. (9.16). Key parameters are the frequency factor f and the penetrability factor
P. The exponent in eq. (9.13) is called the GAMOW factor and yields eq. (9.14).
potential exited levels depend on the shell structure, in particular in terms of overall
spin.
Figures 9.18 and 9.19 show two α-emitting nuclides. In Fig. 9.18 the α-emission
directly and exclusively addresses the ground state of K2. This is illustrated by 212Po,
the last α-transformation step in the 4n + 0 natural transformation chain. Figure 9.19
shows an alternative route: a transformation cascade through several excited
states (226Ra.)
In the context of excited states, it is interesting to note that α-emission (like in β-
transformation) may not only result in the formation of an excited state; it may even
start from an excited state. Figure 9.20 gives an example for 212Bi. In addition to de-
excitation via emission of electromagnetic radiation, some excited states following a
primary β-transformation are subject to a subsequent second primary transformation,
thereby forming a nuclide K3 instead of de-exciting ⊙iK2 to gK2. This nuclide under-
goes α-emission to form stable 208Tl (33.93%) and β−-emission to form unstable 212Po
(64.07%). The ground state of 212Po behaves as illustrated in Fig. 9.18. However, sev-
eral exited states of 212Po are accessible via the β−-transition of 212Bi. A very low per-
centage (compared to the main de-excitation via electromagnetic radiation) de-excite
262 9 α-Emission
212Po
0.3 s 0+
*K1
0% / 8.785 MeV
100%
Q
8.954 MeV
100
208Pb
gK2
gK2 0+
stable
Fig. 9.18: α-emission as direct transformation into the ground state of K2 (212Po ! 208Pb). Kinetic
energy of the emitted α-particle is indicated in the arrow as 8.785 MeV, i.e. less than the Qα-value
by the fraction of the recoil nucleus energy according to eqs. (9.6) and (9.7). Symmetry parameters
are indicated for ✶K1 and gK2. There are no changes in overall spin and parity.
by α-emission directly to 208Pb, i.e. not through ground state 212Po. The energy of
these α-particles is unusually high, even higher than the α-emission from ground
state 212Po to ground state 208Pb.
The α-particle emitted travels a given distance until its kinetic energy approaches
zero. The α-particles typically emitted in a range of 4 to 7 MeV focus their therapeutic
action13 on a short distance of 30 to 80 μm. This causes a high density of ionization
within a short distance. Due to the higher mass of the α-particle relative to β−-particles,
their linear energy transfer (LET) is high. Consequently, ionization effects can be well
focused, resulting in irreparable tumor DNA double-strand breaks close to a tumor
cell. The corresponding LET values for 211At, for example, are 97 keV/μm.14 Several
α-emitters are thus attracting interest in tumor therapy.
Another therapeutic effect in tumor treatment caused by α-particles is achieved
via neutron-induced nuclear reaction on 10B. Once a biological carrier system has
delivered this stable isotope of boron close to the tumor cell (or even inside it, near
226Ra
1600 a *K1 0+
8.65
4K2 3-
4.5
3K2 1-
10.4
2K2 4+
0.96
1K2 2+
1.0
222Rn
gK2
3.823 d 0+
Fig. 9.19: α-emission populating several excited states of K2 (226Ra ! 222Rn). Symmetry
parameters are indicated for ✶K1, ⊙iK2 and gK2. Each arrow shows the logft-values for the
transformation, the corresponding abundance of the separate transformations, and the
corresponding energy of the emitted α-particle. Kinetic energy of the emitted α-particle is
maximum for ✶K1 ! gK2 (4.871 MeV), which is also the most probable transformation (94.04%).
The next most probable one is ✶K1 ! ⊙1K2 (5.95%) with less kinetic energy for the α-particle
(4.684 MeV).
the cell’s nucleus), the nuclear reaction converts the 10B nucleus into an α-particle
and a 7Li nucleus (shown in Fig. 9.21). The kinetic energy of both particles (α and
7
Li) induces ionization along their paths. Again, this leads to double-strand breaks
of the tumor DNA and selective death of the malignant cell. This process is called
Boron Neutron Capture Therapy (BNCT). A commonly used carrier molecule is
10
B-phenylalanin, a modified amino acid that accumulates in proliferating tumor
tissue to a higher degree than in normal tissue.
n + 10 B = α + 7 Li (9:17)
264 9 α-Emission
212Bi
1.009 h *K1 1-
/ 451.2 keV
0% / 1524.8 keV
44% / 739.4 keV
90% / 631.4 keV
64.07%
Q
0.032%
2.252 MeV
4.0
0.6
1.9
1.4
55.3
0
6K2
5K2
4K2 1
3K2 1
2K2
1K2
212gPo gK2 1 0+
0.3 s
0.01106% / 10.755 MeV
208Pb
stable gK3 0+
Fig. 9.20: α-emission starting from excited nuclear states ⊙iK2. Three α-emissions originate (1)
from excited nuclear states of 212Po as populated in initial β−-transformations 212Bi ! 212Po.
Correspondingly, energies of these α-emissions are larger (Eα > 9 and > 10 MeV, respectively) than
the one shown in Fig. 9.18(2) (Eα0 = 8.785 MeV).
Nuclide t½ Production
Ac d U-chain, Th-chain, Ra(p,n)Ac,
Ra . d Th
Ra . d Ac-chain, Th-chain, Ra(n,γ)Ac
Bi . m Ac-chain, Ac/Bi-generator
Bi m Ra-chain, Bi/Pb-Generator
At . h Bi(α,n)At
Tb . h Ta(p,spall), Gd(p,n)Tb
9.8 Example applications 265
NEUTRON
CELL
TUMOR NUCLEUS
CELL
7Li-PARTICLE
Fig. 9.21: Concept of BNCT. A compound containing 10B is irradiated by neutrons. The nuclear
reaction creates two products: an α-particle and a 7Li-particle.
Since MARIE CURIE produced pure samples of 226Ra and other isotopes, long-lived
nuclides present in the naturally occurring transformation chains have not only be-
come an object of research but also a tool for providing α-radiation. The α-particles
were and are utilized e.g. as projectiles for scattering studies on atoms (see RUTHER-
FORD’S irradiations of gold foil) and as projectiles for nuclear reactions (see I. CURIE /
P. JULIOT’S first man-made synthesis of unstable isotopes and the current production
pathways of artificial radionuclides). This will be discussed in Chapter 13.
Furthermore, the emission of α-particles has found application in smoke detec-
tors. The concept is to permanently measure the ionization induced by the interaction
of α-particles with air to register a constant electric current between two electrodes.
The conductivity of air is low, but measurable (<10 pA). If dust or smoke particles are
in the air, they will adsorb the ions formed. Consequently, the electric current is dis-
turbed, which induces an alarm signal (see Fig. 9.22). To guarantee a long shelf-life of
these systems, long-lived α-emitting radionuclides are needed. Typical sources in-
clude 226Ra and 241Am. In order to provide a similar α-emission intensity, a higher ra-
dioactivity is needed for longer-lived nuclides. Thus, typical activities are in the range
of 0.5 MBq and 50 kBq for 226Ra (t½ = 1600 a) and 241Am (t½ = 432.2 a), respectively.
While smoke detectors these days are often constructed without radioactive ma-
terial, there is another application which is rather unknown to most of us: batteries.
Have you ever wondered how satellites and space probes manage to send signals to
266 9 α-Emission
AIR DUST
Fig. 9.22: Concept of using α-emitters in smoke detectors. Pure air provides ionization over the
whole distance, while the presence of dust particles shortens the range of the α-particle
interactions. The electric current is interrupted by dust or smoke.
Earth, even after many years? Remember Pioneer 10, the first man-made object to
pass Jupiter and leave our solar system. Launched in 1972, it communicated with
earth until 2003. How is this possible when solar energy is insufficient to generate
electric power on board?
The answer lies in “radioisotope thermoelectric generators (RTG)”, a type of “nu-
clear battery”. For example, the α-emission of long-lived 238Po (t½ = 87.7 a) interacts
with surrounding matter to create heat, which is transformed into electric power. A
few grams are sufficient to provide several Watts of power. (Alternatively, the β-
emitter 90Y can be used instead.) 238Po was used to generate energy for missions such
as Pioneer 10 and 11 (in deep space) and Viking 1 and 2 (to Mars). Lunokhod, a lunar
rover, utilized 210Po (t½ = 138.4 d) as a power source. In addition to 238Po, 244Cm (t½ =
18.1 a) and 241Am (t½ = 432.2 a) can also be used, depending on the mission duration.
A simpler version of the RTG is the RHU, or “radioisotope heater unit (RHU)”.
These units provide only heat, not electric power, and are used in cold, remote sites
such as outer space, Antarctica and other places on earth.
After samples of 226Ra and other long-lived α-emitters were used as tools to induce
nuclear reactions, a new interest in neutrons arose. Since then, α-sources have
been shown to convert α-particles into high-energy neutrons. Equation (9.18) gives
one of the classical reactions representing an α/9Be-neutron source and Fig. 9.23 il-
lustrates the concept. The nuclear reaction products are stable 12C and the “free”
neutron. The mono-energetic α-particles provided by each α-emitting radionuclide
9.8 Example applications 267
generate mono-energetic neutrons (in the MeV energy region) according to the balance
in mass of 9Be and the kinetic energy of the α-particle emitted from the nuclide.15
α + 9 Be = 12 C + n (9:18)
The α-emitting radionuclides should ideally transform exclusively by α-emission
and should be of long half-life.
The corresponding light-element nuclear reaction partners should have the fol-
lowing properties:
– mono-isotopic (i.e. of 100% natural abundance), to provide a highly efficient α/
n-conversion and to avoid additional nuclear reactions;
– allow for the formation of a stable reaction product nuclide (such as 12C in the
case of 9Be), and;
– have a high nuclear reaction yield.
NEUTRON
C-PRODUC
12C-PRODUCT
Fig. 9.23: Neutron sources based on α-emitting radionuclides. Illustration of the α-induced nuclear
reaction on 9Be, forming a stable nuclear reaction product 12C.
9
Be qualifies as a target for the α/n conversion process. Naturally occurring beryl-
lium consists of 100% 9Be (9Be = “mono-isotopic” among the stable isotopes), it has
a high nuclear cross-section for α-induced nuclear reaction, and it forms stable 12C.
The idea is to dilute the α-source (remember the short range of α-particles) with light chemical
elements, allowing an α-induced nuclear reaction that will form neutrons. However, the α-particle
still loses kinetic energy before entering the 9Be nucleus. Consequently, the kinetic energy of emit-
ted neutrons is often less than the theoretical maximum. Finally, the neutron energy becomes
somewhat continuous.
268 9 α-Emission
Table 9.5 lists four of the most frequently used α-emitters. Depending on the
activity of the nuclides selected, neutron fluxes of about 107 per second are perma-
nently generated. The α/n conversion sources are convenient systems that allow for
basic experiments in nuclear chemistry and are routinely applied in teaching and
training.
16 At that time, radium was considered a miracle cure for all kinds of ailments and aches. Radium
rays were associated with energy and increased performance, so industry used these terms in advertis-
ing and sales-promotions. At that time, crèmes, powders, toothpaste, water, beer, butter, razor blades
and much more simply sold much better if you put radium or radiation in front of the product name.
This ended when the hazards of radioactivity became obvious from the 1920s on, cf. Figure 9.24.
9.9 Outlook
Similar to β-transformations, nuclear shell structure also plays a significant role for
α-transformation, in addition to balances in mass and energy. So far, α-transitions
have been discussed without explicitly considering shell effects. However, Fig. 9.13
already unveiled a difference between α-transitions along (even, even) nuclides for
the 4n + 0 and 4n + 2 natural decay chains and the 4n + 1 and 4n + 3 analogue ones.
For similar Qα-values of ca. 6 MeV, for example, half-lives of the (even, even) nu-
clide chains are orders of magnitude shorter than for the 235U chain.
This effect can be explained by a sort of hindrance for certain transitions. The
idea is to consider α-transformations between (even, even) nuclides as proceeding
“smoothly”, while other (Z, N)-configurations appear somewhat “hindered”, and
thus may only proceed with a sort of “delay”. In addition, their half-lives will be
longer.19 Consequently, a hindrance factor H́ is defined as the ratio between the
For the principal effects of radiation on tissue, see Chapter 2 on “Radiation Dosimetry” in Vol-
ume II of this textbook: Modern Applications of Nuclear- and Radiochemistry.
For further applications of α-particle emitting radionuclide in the context of radiopharmaceuti-
cal chemistry, nuclear medicine and tumor therapy, see Volume II, Chapters 13 on “Life Sciences:
Therapy”.
This qualitatively reminds us of the role of allowed and forbidden transitions for β-transformations,
but hindrance factors discussed in α-transformation processes do not belong to selection rules.
270 9 α-Emission
“real” (experimentally measured) and “expected” (i.e., from (even, even) nuclide α-
transformations) half-lives, as per eq. (9.19).
t1=2 experimental
H́ = log (9:19)
t1=2 expected
More recent models thus refer to correlations of type t½ = f(Eα), which consider nu-
clear shell effects — particularly for (magic) proton numbers Z. One approach uses
Eα-values instead of Qα because according to eq. (9.6) for a large range of A (190 ±
50), there are relatively constant relations of type Eα = 0.973 ± 0.003 Qα, as per eq.
(9.20) (TAAGEPARA, NURMIA 1961).
1 2
logt1=2 = C1 ZK2 Eα − =2 − ZK2 =3 − C2 (9:20)
The point is to look for Z and to introduce constants C1 and C2 reflecting shell effects.
When plotting many of the α-emitting nuclides according to eq. (9.20) in terms of
(ZK2Eα−½ − ZK2⅔) as x-axis vs. log t½ as y-axis, straight lines appear for different sets of
nuclides. When selecting nuclides such as 215Rn, 226Ra, 238U, 232Th, values for C1 and C2
fit to 1.61 (MeV)½ and 28.9, respectively, and create a straight line. This is defined as = 0.
Other nuclides follow different linear trendlines, which are parallel ′ to the zero
hindrance line towards longer half-lives. Hindrance factors of 0 < H́< 1.2 indicate
minor hindrance (reflecting nucleon numbers far from magic), while larger values
of H́correspond to nuclides with nucleon configurations close to magic numbers.
This may be further structured by another approach (DASGUPTA-SCHUBERT, REYES,
2007) according to the four possible combinations of nucleons, i.e. (even, even), (even,
odd), (odd, even) and (odd, odd) (as per eq. (9.21); the coefficients a, b and c are listed
below).20
1 1
1
ðt1=2 in secondsÞ
logt1=2 = a + b A =6 ZK1 =2 + c Qα Z − =2 (9:21)
(Z,N)✶K a b c
For more details, see: A Vértes, S Nagy, Z Klencsár, RG Lovas, F Rösch (eds.), Handbook of Nu-
clear Chemistry, second edition, Springer 2010, Volume I, Chapter 2: Basic properties of the atomic
nucleus, T Fynes.
9.9 Outlook 271
NEUTRON
SMALLER PARTICLE
Fig. 9.25: Directions for emission of clusters smaller or larger than the α-particle. (left) Either one
neutron or one proton, or (right) a 14C cluster, for example.
Ac 225Ac
Fr 221Fr
1
Rn 218Rn
R
At 2
Po
Bi
208Pb
Pb
Tl
Fig. 9.26: 14C cluster emission from 222Ra shown within the Chart of Nuclides. The Chart of Nuclides
identifies nuclides known to emit clusters with a gray triangle at the upper right corner of the
nuclide’s square. Note that the dominant pathway for 222Ra to stabilize is via α-emission to 218Rn
(1). 14C-emission from 222Ra yields double magic 208Pb (2). Only one out of about 1010 222Ra atoms
utilizes 14C-emission instead of α-emission.
option for A > 220 and Z > 87 nuclides. Nevertheless, pre-formation of such large
clusters in heavy nuclei is very improbable. Consequently, only the creation of
transformation product nuclides with stabilized nucleon shells increases the over-
all probability of cluster transformation.
✶
ZK K Cluster ZK ΔA K
Fr 14
6 C8 – Ti double
Ra Pb magic
Ra Pb magic Z
Ra Pb magic Z
Ac Bi
Th 18
8 C10 – Pb double
magic
Th 24
10 Ne14 – Hg
Pa Bi
U Po
Figure 9.28 identifies the area of p-emitting and n-emitting nuclides within the
Chart of Nuclides. Proton-emitting nuclides are very close to the proton drip line of
the Chart of Nuclides. There are several such nuclides; however, all have very short
half-lives of microseconds to nanoseconds (for most of these, the precise value is
not yet determined). Only a few nuclides around the lanthanide group show half-
lives of milliseconds, such as 151Lu transforming to 150Yb with t½ ≈ 81 ms.
Neutron emission is a process that happens mainly in the context of stabiliza-
tion of extremely neutron-rich “fragments” created within fission processes. The
nuclides involved here are of mass numbers around A = 100 and 140, which parallel
the frequent mass numbers of fission processes (see Chapter 10).
These transformations do not require “pre-formation” of a cluster such as the α-particle. Also
(in the case of proton emission), the coulomb barrier is lower compared to α-emission.
274 9 α-Emission
Fig. 9.27: Pathways of proton and neutron emission compared to β-transformations, α-emission
and cluster emission. Proton emission (1) is an alternative to β+- or electron capture transformation
(2) of proton-rich nuclides, while neutron emission (3) is an alternative to β−-transformation (4) of
neutron-rich nuclides. While β-transformation follows the isobar line, emission of protons
proceeds along isotones and emission of a neutron takes place along isotopes. The dominant
α-emission is shown with (5) along an isodiaphere relative to the much less abundant emission of
carbon or larger clusters (6).
9.9 Outlook 275
114
126
50
82
EMISSION OF SINGLE NEUTRON
28
instead of n → p conversion
20
50
8
2 20 28
2 8
Fig. 9.28: Area of extreme p-rich and n-rich nuclides within the Chart of Nuclides undergoing p- or
n-emission, respectively.
10 Spontaneous fission
Aim: In addition to β-processes and α-emission, there is a third option for naturally
occurring unstable nuclides to stabilize: spontaneous fission. It becomes an option
for very heavy unstable nuclides around A ≥ 232 and Z ≥ 90 (starting from 232Th, 235U
and 238U). Typically, it accompanies α-emission. For unstable nuclides of A > 250,
the fraction of spontaneous fission relative to α-emission increases. For many of the
artificially produced transactinide nuclides, spontaneous fission becomes the domi-
nant route of transformation.
Fission is, by definition, the division of a nucleus into two or more parts with
intermediate mass fragments, usually accompanied by the emission of neutrons,
gamma radiation and (rarely) small charged nuclear fragments. Consequently,
there is a completely different mechanism involved in the fission process com-
pared to β-transformation and α-emission.
The most notheworthy impact of “fission” arises from “induced fission”, repre-
senting a giant source of nuclear energy. Induced fission, however, will be dis-
cussed in Chapter 13 on “nuclear reactions”. Nevertheless, many features of the
fission of unstable heavy nuclides apply to both the spontaneous and the induced
variant, and are introduced in this chapter.
This chapter discusses:
(i) the formation of primary fission fragments,
(ii) the fate of these fragments,
(iii) some of the principal theoretical models and mathematical explanations of fission.
10.1 Introduction
Spontaneous fission (sf) represents one more option for naturally occurring unsta-
ble nuclides to stabilize – in addition to the β-processes and to α- and cluster emis-
sion. The number of unstable nuclides “naturally” transforming via the fission
process is much lower compared to the two other routes. Spontaneous fission is
most relevant from an academic point of view and for research on super-heavy
chemical elements; its practical usage is quite limited. The rather dramatic impact
of “fission” arises from “induced fission”, representing an enormous source of nu-
clear energy. Induced fission, however, belongs to the chapter “nuclear reactions”.
Nevertheless, many features of the fission of unstable heavy nuclides apply to both
the spontaneous and the induced variant.
What does “fission” mean? Fission is the concept of dividing a very large nucleus
✶
K1 into two medium-large fragments K2 and K3 (via “binary fission”, eq. (10.1)), or,
more rarely, into three fragments K2, K3 and K4 (via “ternary fission”, eq. (10.2)). The
process is exothermic and the transformation energy gained is ΔE = Qsf.
https://doi.org/10.1515/9783110742725-010
278 10 Spontaneous fission
✶
K1 ! K2 + K3 + ΔE (10:1)
✶
K1 ! K2 + K3 + K4 + Δ E (10:2)
The process of fission, however, is not “just” the division of a certain large nucleus.
It involves (a) several steps prior to the moment fission occurs and (b) interesting
successive follow-up steps, until the initial fission fragments have finally turned
into the individual fission product nuclides. During these successive transforma-
tions, neutrons, electromagnetic radiation, β−-electrons and electron neutrinos are
also emitted. The process of nuclear fission is thus much more complex than that of
the β- and α-transformation processes. In the following, spontaneous fission is ex-
plained with the focus on binary asymmetric fission of the uranium isotope 238U.
These nuclides are the parents of the three naturally occurring transformation chains.
Spontaneous nuclear fission was discovered by Soviet physicists PETRZHAK and FLEROV in 1940,
carefully registering transformation events in uranium compounds. This was slightly later than the
discovery of induced fission in 1938 by HAHN, STRASSMANN and MEITNER.
10.2 Occurrence of spontaneous fission 279
transactinides
126
Fig. 10.1: Occurrence of nuclides within the Chart of Nuclides partially undergoing spontaneous
fission.
The relations in mass between the initial nuclide and the fission fragments are differ-
ent compared to α-emission. Suppose α-emission is written as ✶K1 ! K2 + K3 + ΔE,
where K2 represents the new nuclide and K3 the α-particle. The mass ratio of K2:K3
here is typically (>200):4, i.e. >50:1. In contrast, spontaneous fission forms K2 and K3
as two separate medium-large nuclei with masses of comparable order of magnitude.
For a symmetrical division into two fragments of identical mass, the ratio between
the product fragment masses would be 1:1. In this case, balances would ideally be
A
K2 ≈ AK3 ≈ ½ AK1. Indeed, this happens for very heavy nuclides of Z ≥ 100. For iso-
topes of thorium and uranium (and other nuclides), however, binary fission is asym-
metric. In most cases, a larger (K2) and a smaller (K3) fragment are formed with
varying mass distributions. For 238U the two different primary fission fragments have
masses of about 138 ± 15 and about 100 ± 15, i.e. at a ratio of ca. 1.4:1. Figure 10.2 sche-
matically illustrates the difference between α-emission and fission concerning distri-
butions of mass. Figure 10.3 illustrates the distribution of mass numbers between the
two different primary fission fragments for 238U. The most relevant fragments created
have proton numbers of Z2K2 + Z1−Z2K3 = Z1K1, with Z2K2 and Z3K3 ranging from ca. 47 to
70 and 31 to 51, respectively. Primary fission fragments represent much lighter chemi-
cal elements compared to the initial nuclide.
280 10 Spontaneous fission
α-EMISSION A=4
A = 238 A = 234
A = 238
Fig. 10.2: Mass scale of unstable nuclides undergoing α-emission and spontaneous fission. 238U is
chosen as an example. Radii are proportional to A⅓, as are the dimensions of the circles. The
nucleus undergoing spontaneous fission is depicted as a sphere. Note that this is not correct,
because spontaneous division of a large nuclide requires a nonspherical, i.e. deformed, nucleus.
120 80
238
118 158
148 90
Fig. 10.3: Spontaneous fission of 238U yields asymmetric mass distributions between the two primary
fragments with combinations of AK1 = AK2 + AK3. The most commonly occurring pairs are AK2 + AK3 ≈
138 + ≈100, occurring in about 8% of all sf events. When the mass number of A2 shifts to larger (e.g.
145 or 150) or lower (e.g. 130 or 125) values, the value of A3 adjusts correspondingly. The probability
of forming these pairs, however, decreases, as indicated by the thickness of the lines.
10.2 Occurrence of spontaneous fission 281
248Fm
36 s
248Cm
7.87, 7.83, … , sf,
3.40+5 a
235U 238U
0.7204 99.274
26 min 7.038+8 a 298 ns 4.468+9 a
232Th l (0.07) 4.398 …, sf l 2514, … 4.198 …, sf
e- , e- sf 2 , , e-
100
7.038+8 a
4.013, 3.950, …, sf
, e-
Fig. 10.4: Examples of nuclides simultaneously undergoing spontaneous fission and other types of
transformation. The different colors indicate the various transformation routes: spontaneous
fission = green, α-emission = yellow, β−-emission = blue, β+-emission and electron capture ε = red,
cluster emission = gray. White fields refer to metastable isomers to be discussed in Chapter 11. The
size of each triangular area symbolizes the relative abundance of that pathway.
282 10 Spontaneous fission
10.3.1 Deformation
Why should a nucleus divide? According to the liquid drop model, strong forces
keep nucleons closely together as an ensemble, and there is a mass defect, i.e. a
gain in energy, compared to the same number of unbound nucleons. The only force
which can “disturb” this equilibrium is coulomb repulsion between the protons.
The LDM describes protons as being homogeneously distributed within the nucleus.
But what if Z1 protons were grouped into two (somehow) locally separated semi-
centers with proton numbers Z2 and Z3 (Z2 + Z3 = Z1)? Something must happen prior
to this moment: deformation. Remember the concept of a “liquid drop”. How does a
liquid drop behave? It may change its shape by switching from spherical to oblate
or prolate. Immediately, there are several consequences, listed in Tab. 10.1.
Surface and surface Compared to a sphere, a deformed sphere of identical volume shows a
tension larger surface area. The surface tension also increases.
Shell structures According to the shell model, the shells of deformed nuclei obtain are
rearranged in terms of energy and occupancy, as per Fig. ..
Mean nucleon binding New nucleon arrangements within the shells of a deformed nucleus may
energy improve the mean nucleon binding energy and stabilize the deformed
nucleus in a new ground state, due to a local minimum in potential energy.
The general theories describing the stability of a nucleus of an atom are the liquid
drop and the nuclear shell model.3 According to the WEIZSÄCKER equation, surface
energy and coulomb repulsion play a role. For a sphere S, the following character-
istics are also known:
– volume is SV = 4/3 πr3,
– the relationship between mass number and volume and size is SAV ≈ r⅓
– surface area is SS = 4πr2,
– the relationship between the mass of a nucleus (reflecting the overall number
of nucleons) and the surface and radius of a sphere is SAS ≈ r⅔.
In the context of the LDM, nuclei may transform (deform) into an ellipsoid. For
an “ellipsoid of rotation”, the three individual radii a, b and c turn into a = rlong
and b = c = rshort (see Fig. 10.5). For a rotational prolate ellipsoid E, the volume is
E
V = 4π/3 (rlong)⋅(rshort)2, as per eq. (10.3). The surface of an ellipsoid is expressed
through eq. (10.4).
minor axis
rshort
major axis
rlong
Fig. 10.5: Geometry of a prolate ellipsoid of rotation with two axes of rlong and rshort.
These two models are applied to unstable nuclei as well, but in particular for very heavy nuclei
some of the essential inputs in both models may not apply – at least in detail.
284 10 Spontaneous fission
Suppose this ellipsoid has the same volume as a sphere containing the same
number of nucleons; then the surface of the two bodies (as per eq. (10.4)) will differ
depending on the ratio between the two axes of the rotational ellipsoid. The differ-
ence increases the longer the deformation continues.4
4 4
E
V= π ðabcÞ = π rlong ðrshort Þ2 (10:3)
3 3
" #0.625
ðabÞ1.6 + ðacÞ1.6 + ðbcÞ1.6
E
S ≈ 4π (10:4)
3
The radius r along an axis of rotation becomes smaller for prolate deformation and
larger for oblate deformation. The increasing ratio of rlomg/rshort from 1 (spherical) to
>1 (prolate) or <1 (oblate) (as per Fig. 10.6) is described by the degree of eccentricity ε.
This may be expressed by a deformation parameter εDEFORMATION, which involves the
changes in radii r between the deformed shape and the spherical one, following eq.
(10.5). Here ro stands for the radius of the sphere and Δr is the difference between
long and short axis of the rotational ellipsoid, Δr = r(long axis) − r(short axis).
rðlong axisÞ − rðshort axisÞ
ε DEFORMATION ≈ (10:5)
r0
For rlong:rshort = 2:1 or 4:1, the surface of a prolate ellipsoid relative to that of the sphere increases
by factors of 1.077 and 1.279, respectively. For a rotational oblate ellipsoid (rlong:rshort = 1:2 or 1:4), it
is even more extreme: 1.095 and 1.428, respectively.
The corresponding coulomb term of the WEIZSÄCKER equation was − γLDM Z2 / A⅓.
10.3 Pathways of fission 285
Fig. 10.6: Ellipsoids of rotation along one selected axis, forming either prolate or oblate nuclei
depending on the ratio between rlong and rshort.
Surface energy ESURFACE is a force keeping a “liquid drop” body of nucleons together.6
Increasing deformation may be expressed by a quadrupole stretching factor qsα. Upon
deformation, the surface energy ΔESURFACE is positively affected according to eq. (10.8),
while for coulomb energy ΔECOULOMB the opposite occurs, as per eq. (10.9). The changes
in surface and coulomb energy when deforming from a spherical to an ellipsoidal nu-
cleus are expressed through eqs. (10.10) and (10.11).
Similar to the surface tension of a drop of water or even more extreme: mercury.
286 10 Spontaneous fission
2 qs 2
E SURFACE
E = E 1+
S SURFACE
α (10:10)
5
1 qs 2
= E
E COULOMB S COULOMB
E 1+ α (10:11)
5
While these equations give a relationship between energies of deformed nuclei rela-
tive to the spherical one, individual energies can also be calculated directly following
the WEIZSÄCKER equation. Here, analogue coefficients (in units of MeV) for surface
and coulomb energy are denoted by αSURFACE and αCOULOMB. For surface energies, the
radius ro and the mass number A are linked with a surface tension factor STENSION,
eq. (10.12). In the case of coulomb forces, the number of protons appears in addition
to values of ro and A and the electric charge unit e, eq. (10.13). Table 10.2 lists the
resulting expressions for the ratio between the changes in the two energies.
2 2
S SURFACE
E = 4π r2O STENSION A =3 = α SURFACE A =3 (10:12)
3 Z2 e 2 Z
S COULOMB
E = 1
= αCOULOMB 1 (10:13)
=
5 r0 A 3 A =3
Tab. 10.2: Changes in coulomb and surface energies for deformed nuclei. Individual equations and
resulting equivalences.
Z2 (.)
≈
A
SURFACE COULOMB
ΔE = ΔE (.)
S SURFACE (.)
E 2
=
S ECOULOMB 1
There will exist a degree of stretching for which the stabilizing effect of deforma-
tion on surface energy ΔESURFACE is identical to its negative effect on coulomb en-
ergy ΔECOULOMB, eq. (10.15). Here, a quotient of type Z2/A appears, eq. (10.14),
indicating that the density of protons per volume (reflected by A) plays a signifi-
cant role. Table 10.3 list values of Z2/A for select nuclides transforming (at least in
part) by spontaneous fission. According to eqs. (10.8)–(10.11), this corresponds to
S SURFACE S COULOMB
E / E = 2/1, eq. (10.16).
10.3 Pathways of fission 287
Equations (10.15) and (10.16) express an equilibrium between the changes in coulomb
energy and the changes in surface energy for a ratio of SESURFACE / SECOULOMB = 2.
When the ratio between the two contributions is changing, the value x is called the
fissility parameter, shown in eqs. (10.17)–(10.19).
S COULOMB
E
=x (10:17)
2S ESURFACE
α COULOMB Z2
x= (10:18)
2α SURFACE A
2
Z =A
x= critical (10:19)
Z =A
2
Equation (10.18) involves basic parameters of nuclides, namely the mass number A,
proton number Z and the two coefficients αC and αS. The value of x is thus directly
derived.7 The interesting point is to define a “critical” value for Z2/A of a given nu-
cleus. Given αC / 2αS equals the reciprocal of Z2/A, i.e. αC / 2αS = (Z2/A)−1, eq. (10.18)
turns into (10.19). This term is called (Z2/A)critical. Thus, it is the proton number Z
per mass number A of a nucleus, somehow a “proton density” per volume, which is
responsible for an eventual fission process. For fissile nuclides, x < 1. For 235U, 238U
and 252Cf, for example, the values are 0.749, 0.740 and 0.793, respectively.
At a transition state (or so-called saddle point8), the nucleus may switch to a new
geometry by forming a neck between two semi-centers of the ellipsoid, shown in
Fig. 10.7. From now on, coulomb repulsion may take over the process in an irrevers-
ible way, and within a very short period of time the phase of division starts. This
moment is called the scission point.
With αC = βLDM = 15.94 MeV and αS = γLDM = 0.665 MeV, the quotient αC / 2αS gives a proportional-
ity factor of 0.0203, and the fissility parameter x is now directly calculated by simply introducing
specific numbers for Z and A of a candidate nucleus.
. . . which is about 10−20 s prior to the fission event, see below.
288 10 Spontaneous fission
Fig. 10.7: Deformation (2) of a large spherical nucleus (1) to prolate shape (3) and subsequent necking
(5) of the deformed nucleus. At a saddle point (4), coulomb repulsion between the two internal centers
of high proton charges of the two fragments may succeed over surface tension, and at a scission point
(6), the process may continue towards division to form two separate (primary) fragments.
The energetics of spontaneous fission according to the LDM explain the general aspects
of this type of transformation. As already seen for β- and α-transformations, nucleon
shell models overlay the LDM. The stability (nucleon binding energy) is enhanced for
nuclei containing paired nucleons and magic nucleon numbers. Simply speaking,
whenever there is a chance to collect a certain number of nucleons, why not prefer
those reflecting magic numbers? And in fact, this is the strategy! What about the re-
maining nucleons? They would form the neck, as shown in Fig. 10.8. When fission
takes place, the neck ruptures. The neck position is statistically determined, and the
two parts of the neck (belonging to either of the two fragments) combine with the cor-
responding fragment. This will add a certain number of protons and neutrons to the
two “pre-formed” fragments. Once the neck has divided, two separate fragments of
high proton number (Z ≥ 28 or ≥ 50) are repelled through coulomb forces.
For the example of the heavy fragment AK2, because the magic numbers involved
are Z = 50 and/or N = 82, this would ideally correspond to a fragment of Z = 50 plus
N = 82 yielding A = 132, with a double magic nucleon number. In parallel, nucleon
numbers may alter as the corresponding “piece” of the neck is added, though at
10.3 Pathways of fission 289
Z = 28 Z = 28 Z = 28+x
N = 50 N = 50 N = 50+y
1 2 3
Z = 50 Z = 50 Z = 50+x
N = 82 N = 82 N = 82
82+y
Fig. 10.8: The fate of the neck in the fission mechanism. A deformed nucleus starting to accumulate
double magic nucleon numbers into two semi-centers of Z = 28 + N = 50 and Z = 50 + N = 82, while
the remaining nucleons form the neck between them (1). The neck divides in a statistical pattern, and
the two remaining parts of the neck combine with the corresponding fragment to add a certain
number of x protons and y neutrons to the fragment (2). Once the neck has divided, two separate
fragments of high proton number (Z ≥ 28 or ≥ 50) remain and are repelled through coulomb forces (3).
least one sort of nucleon would try to “conserve” its magic value. This would ex-
plain a mass distribution of the larger fragment of Z = 50 (fixed) and N = 79, 78, etc.
or N = 81, 82, etc., but also N = 82 (fixed) and Z = 51, 50, etc. or Z = 53, 54, etc. The
distribution of the resulting primary fission fragments thus shows a high abun-
dance of certain combination of mass numbers of the two fragments, while other
combinations are less abundant. However, the sum of the nucleons within the ini-
tial fission step remains constant. Figure 10.9 gives the profile of mass distributions
of primary fragments in the spontaneous fission of 252Cf.
The effect of nucleons existing pairwise also improves overall nucleon binding en-
ergy. This holds true both for neutrons and protons. Consequently, even (versus
odd) proton or neutron mass numbers should be more common in the initial fission
290 10 Spontaneous fission
7
252 Cf
spontaneous fission yields for 252 Cf (%)
6
0
60 80 100 120 140 160 180
fragment distribution. This is seen in Fig. 10.9 and is particularly obvious when the
number of protons is considered, as in Fig. 10.10.
With increasing mass of the nuclides and with increasing neutron number among
isotopes of the same element, spontaneous fission tends to turn from asymmetric
into symmetric fission. Here, the two primarily formed fission fragments are of simi-
lar or identical mass number. This means that the initial nucleus divides into two
more-or-less equal fragments of AK2 ≈ AK3 ≈ ½AK1, as per Fig. 10.11. Figure 10.12
Note that for sf of 252Cf, the yields of individual fragment masses are given in percent relative to
the fraction of 252Cf undergoing sf. However, the sf branch of 252Cf is just 3% relative to the domi-
nant α-transformation pathway.
10.3 Pathways of fission 291
0.60
54 56 58 60 62 64 252 Cf
0.50
spontaneous fission yields for252Cf (%)
0.40
0.30
0.20
0.10
0.00
50 55 60 65 70
primary fission fragments proton number Z
Fig. 10.10: Distribution of proton number Z of the two primary fission fragments in the spontaneous
fission of 252Cf.
ASYMMETRIC SYMMETRIC
Z Cf Es Fm Md No Lr Rf
N
1.1 s
146
36 s
148
2.3 s
150
13.08 a 55 s 6.2 ms
152
2.645 a 20.47 d 3.24 h 2.91 s 12 ms
154
100.5 d
157
158
14
1.5 s 31.8 d
% Yield
10
159
6
2
5 ms
160
110 150 110 150 110 150 110 150 110 150 110 150 110 150
A of fragments 2 3
Fig. 10.12: Asymmetric and symmetric mass distributions of initially formed fragments in binary
spontaneous fission. Isotopes of the heavy actinide elements californium, einsteinium, fermium,
mendelevium, nobelium and lawrencium and the first transactinide element rutherfordium are
arranged according to proton and neutron numbers of the fissioning nuclides. Half-lives of ground
state isomers are added in the corresponding boxes. Spontaneous fission becomes more
symmetric (1) for increasing neutron number at constant proton number, (2) for increasing proton
number at constant neutron number, and (3) with increasing mass number (from: DC Hoffmann and
MR Lane, Radiochimica Acta 70/71 (1995) 135).
shows the occurrence of symmetric and asymmetric fission profiles for nuclides of
the very heavy elements of Z = 94 – 104.
The neck model explains this trend in mass distribution. For less heavy nuclides the
two fragments attract magic compositions of type (Z = 50, N = 82) and (Z = 28, N = 50).
10.3 Pathways of fission 293
Increasing the nucleon numbers does not change this pattern; thus, it is the neck that
increases in mass and size. However, as even more nucleons become available (starting
at element Z = 100), a mass distribution of type two times Z = 50 becomes possible. This
turns into binary symmetric fission. At the same time, the neck shrinks, because the two
fragments accumulate almost all the nucleons available to obtain shell occupancies ap-
proaching magic numbers.
This can be summarized by three general tendencies. Asymmetric fission be-
comes more symmetric:
1. for increasing neutron number at constant proton number,
2. for increasing proton number at constant neutron number,
3. with increasing mass number from A = 246 (100Fm146) or 250 (98Cf152) to A = 262
(102No160 or 104Rf158).
The mechanism of spontaneous fission discussed so far refers to the formation of two
fission fragments. However, “ternary” fission10 also occurs in parallel. “Ternary” is
the division of the initial nucleus into three primary fragments, as per Fig. 10.13. Me-
chanically, this refers to the ternary fission of a neck that splits in two places instead
of one. Consequently, a minor component remains (let us denote it as K4), which is
not attributed to the two essential primary fragments, following eq. (10.2). There are
two questions: What is the new fragment AK4? And what is the ratio between binary
and ternary fission?
Fragment distribution in ternary spontaneous fission: For the new primary frag-
ment, one would expect a statistical distribution of mass numbers theoretically
ranging from 1 (the proton or the neutron) to ≤ 40, because the pool of nucleons
available is limited to those found in the neck. Whenever the impact of the shell
model holds true, this spectrum of fragments should represent a few fragments of
particular stability. Indeed, the majority of the lightest fragments are α-particles
(generated in about 90% of all events). It may vary down to fragments of hydrogen
or up to fragments of carbon or even argon. The mass distribution of the two larger
primary fragments is thus shifted to somewhat lower values. When the binary frag-
ments are, for example, binaryAK2 ≈ 135 ± 15 and binaryAK3 ≈ 95 ± 15, then the ternary
fragments are ternaryAK2 ≈ binaryAK2 – 4 or ternaryAK3 ≈ binaryAK3 − 4. This is shifted in a
few cases, according to the somewhat smaller or larger mass number of ternaryAK4.
Experiments have demonstrated that this small fragment is released perpendicular
to the direction of the two large primary fragments (thereby proving the postulated
BINARY TERNARY
Fig. 10.13: Illustration of the asymmetric binary and ternary divisions of a fissile nucleus.
mechanism of ternary fission). While the recoil energies of fragments K2 and K3 are
similar to those for binary fission, the kinetic energy of each α-particle released as
third primary fragment in ternary fission is large, reaching about 16 MeV.11
10.4 Energetics
10.4.1 Trends
The motivation to transform via spontaneous fission is the same as for any other pro-
cess of exothermic radioactive transformation, namely to increase the mean nucleon
binding energy. Along the curve of mean nucleon binding energy according to the
WEIZSÄCKER equation, any shift from a nucleus of very high mass number A to inter-
mediate mass nuclei increases the mean nucleon binding energy, as per Fig. 6.7. In
this context, fission is similar in principle to α-emission, but the gain in mean nu-
cleon binding energy is much higher than in the case of ΔA = 4.
This is about 10 MeV larger than observed for α-particles released during α-transformation of
unstable nuclides.
10.4 Energetics 295
Fission of heavy nuclei into two smaller fragments belongs to a process of utmost
importance as a source of energy. The absolute value of the energy of the spontane-
ous fission process follows eq. (10.1) for a binary asymmetric spontaneous fission.
Equations (10.20) and (10.21) utilize real masses or mass defect or mass excess val-
ues, respectively:
Each of the two primary fission fragments shows an increase in ĒB relative to the
very heavy initial nucleus. Compared to α-emission, the gain in energy in terms of
ΔĒB = ĒB(K2) − ĒB(✶K1) and ΔĒB = ĒB(K3) − ĒB(✶K1) is much higher.
For α-emission, let us consider an A = 238 nuclide emitting an α-particle. ĒB val-
ues are ≈7.570 MeV for 238U and ≈7.597 MeV for 234Th, yielding ΔĒB ≈ 0.047 MeV. For
fission, let us again take the same nuclide of A = 238. In the hypothetical case of
symmetric fission, for each of the A = 119 product nuclei ĒB is ca. 8.35 ± 0.10 MeV.12
The difference is ΔĒB ≈ 0.78 ± 0.10 MeV, which is larger than in the case of α-emission
by a factor of almost 20.
When calculating the overall (instead of mean) nucleon binding energy, each
of the 238 nucleons thus improves by ca. 0.8 MeV, yielding an overall nucleon bind-
ing energy of ca. 190 MeV per one nuclide of A = 238! It is among the largest sources
of energy in the universe and the essence of “nuclear power”.
Simple mathematics would state that the energy gained by dividing one large nu-
cleus into two identical pieces should be the same as that needed to recombine
them into the original nucleus.
According to eqs. (9.11) and (10.6), ECOULOMB energy for 238U and the two hypo-
thetical primary fragments of A/2 = 119 and Z/2 = 46 is about 197 MeV. This is about
7 MeV larger than the energy released in the fission process. This value is called the
“fission barrier” Hfb, described in eq. (10.22). The fission barrier is the difference in
energy between the corresponding nuclear state of the nucleus prior to fission and
There are many unstable nuclides of A = 119 formed along the isobar . . . .
296 10 Spontaneous fission
the maximum potential energy needed to fission.13 Hence, Hfb is >0 (and typically
about 5–8 MeV), which would make spontaneous fission energetically impossible.
These strongly deformed and extremely neutron-rich primary fragments observe ex-
treme coulomb repulsion, thereby obtaining a defined kinetic energy and a high ve-
locity. “Flying” away at that velocity, the fragments release single excess neutrons
(about 2–3 so called “prompt” neutrons, see below) and transfer kinetic energy.15
The height of the fission barrier depends on several parameters and is significantly determined
by the grade of deformation, the size of the neck, and the kind of fission (symmetric or asymmetric).
Here comes the relevance of induced fission. Whatever tool is employed to transfer energy (par-
ticles with kinetic energies, electromagnetic radiation etc.) to a potential candidate nucleus for fis-
sion, this imported energy may help to overcome the fission barrier. Consequently, the fission
process will become possible or (if it was already allowed through the tunneling effect) will be facil-
itated significantly; see Chapter 13 on nuclear reactions.
Maximum neutron kinetic energies En are about 10 MeV, corresponding to a velocity of ca.
107 m/s (which is about 1/30 of the speed of light in vacuum). Due to spatial distributions, adsorp-
tion and interaction, only a few neutrons are released with the maximum kinetic energy. Most of
the emitted neutrons show kinetic energies of ca. 0.2 to 2 MeV.
10.5 Velocities of spontaneous fission 297
SADDLE POINT
potential energy U (MeV)
≈ 5-8 MeV
TUNNELING
3
1
GROUND STATE
SCISSION POINT
degree of deformation
Fig. 10.14: Change of potential energy of the nucleus depending on the degree of deformation. The
nucleus moves from the ground state of the deformed nucleus (1) either to surpass the fission
barrier (2) at the saddle point and further to the scission point or by quantum mechanics tunneling
(3) through the barrier.
actinide elements, for example, are 1.405⋅1010 years for 232Th and 7.038⋅108 years
for 235U, isotopes of heavy actinide or of transactinide elements (Z ≥ 104)16 show
half-lives in the range of seconds and fractions of a second, given in Tab. 10.3.
Spontaneous fission is a dominant transformation pathway for super-heavy nuclides, and val-
ues of t½sf approach microseconds for the heaviest elements identified so far. For details, see Chap-
ter 10 of Volume II on the “Chemistry of transactinides”.
298 10 Spontaneous fission
Figure 10.15 shows the half-lives of many nuclides undergoing spontaneous fis-
sion in terms of the quotient Z2/A. One may further discriminate between nuclides
of even or odd nucleon numbers (or even or odd mass numbers). For similar values
of Z2/A, even nucleon number configurations result in much shorter half-lives by
many orders of magnitude. For one and the same heavy chemical element (Z = con-
stant), half-lives of the isotopes change depending on the Z2/A value. The larger the
abundance of protons per overall nucleon number is, the faster the spontaneous
fission proceeds. This is illustrated in Fig. 10.16 for even proton number elements.
Nuclides of increasing mass number have an excess of neutrons over protons. If the
excess of neutrons is expressed as the ratio N/Z, the WEIZSÄCKER equation shows val-
ues of N/Z = 1.55 ± 0.05 for heavy nuclides of proton number Z = 90–100 and mass
10.6 Follow-up processes of initial fission 299
1,E+20
(e,e)
1,E+16
(o,o)
(e,o),(o,e)
1,E+12
1,E+08
t½ (s)
1,E+04
1,E+00
1,E-04
1,E-08
34 35 36 37 38 39 40 41 42 43 44 45 46
Z²/A
Fig. 10.15: Half-lives of nuclides undergoing spontaneous fission depending on the quotient Z2/A.
(Even, even), (odd, odd) and (even, odd) or (odd, even) nucleon compositions are identified.
1,E+18
90 92
1,E+16
94
1,E+14
1,E+12 96
98
1,E+10
1,E+08
t½ (s)
1,E+06
100
1,E+04 104
102
1,E+02
106
11,E+00
E+00
1,E-02
1,E-04
1,E-06
33 34 35 36 37 38 39 40 41 42 43 44
Z2/A
Fig. 10.16: Half-lives of nuclides undergoing spontaneous fission depending on the quotient Z2/A along
a constant proton number Z, representing isotopes of one and the same element. Z = 90 (Th), Z = 92 (U),
Z = 94 (Pu), Z = 96 (Cm), red line, Z = 98 (Cf), Z = 100 (Fm), Z = 102 (No), Z = 104 (Rf), Z = 106 (Sg).
300 10 Spontaneous fission
numbers of A = 230–250. For the two initial fragments of the spontaneous fission of
238
U, for example of A = 138 and A = 100, these ratios are ≈1.47 and ≈1.34 only, respec-
tively.17 However, the large excess of neutrons characteristic for the very heavy
nuclide is distributed between the two primary fragments. Consequently, the two
primary fission fragments A2K2 and A3K3 also have an excess of neutrons much
too large for their intermediate mass. Primary fission fragments are thus not en-
sembles of nucleons arranged to satisfy the LDM or NSM. The terminology “fragment”
rather than “nucleus” or “nuclide” is preferred for this reason and approaches are
required to get rid of the excess of neutrons. There are two approaches for handling
the excess in the primary fission fragments, namely separation of neutrons or succes-
sive transformation of a neutron into a proton.
Secondary fission fragments do not yet exist at their nuclear ground state, but in
highly enriched nuclear states instead. Such a fragment releases a significant part
of its energy directly by emission of energy in the form of electromagnetic radiation,
namely γ-rays (see Chapter 11). This finally yields states of nuclear structure that
can be discussed in the framework of the LDM and NSM. The species formed after
the separation of neutrons and of electromagnetic radiation from fission fragments
are now called primary fission products.
Tab. 10.4: Properties of initially formed fission fragments and the subsequent processes.
Have a neutron excess and are Relax a neutron Lowering mass Prompt neutrons
(according to their mass number A and by neutron number A
excess of neutrons over protons) separation
far from the optimum nucleon binding
Eject a neutron Lowering mass Delayed neutron
energy of the corresponding isobar,
whenever Qβ > δĒB of number A
leading them to . . .
the “last” neutron
The primary fission products created are now well characterized by nuclide termi-
nology, namely mass number, proton number and neutron number. They have nu-
cleon shell occupancies as described by the nuclear shell model. But these products
are still neutron-rich. Because they are of mass number <200, further stabilization
occurs via β−-transformation. The transformation optimizes the ratio of neutrons-to-
protons, thereby increasing the mean nucleon binding energy along an isobar. It
occurs at the left side of the β-transformation isobar parabola according to Fig. 6.11
(see Fig. 10.17).
This happens if the binding energy of the “last” neutron is less than the Qβ-value of the corre-
sponding n ! p + β− + ν̅e pathway, as per eq. (9.21).
This effect becomes particularly relevant in the context of induced nuclear fission, as it is the
tool to actively control the regime of a nuclear reactor – see Chapter 11 of Volume II.
302 10 Spontaneous fission
100 1
EB Z
80
nts
60 g me sion
ra is
yf cf
m ar e tri 2
i
pr m
ym
40 n as
i
5 4 3
20
A= const.
0
0 20 40 60 80 100 120 140 N
Fig. 10.17: Excess of neutrons in primary fission product leads to subsequent β−-transformation
processes. (1) A heavy nuclide divides, forming large (2) and small (3) primary products. These
have an excess of neutrons and are more or less far away from the line of β-stability (4). The excess
of neutrons over protons is managed by converting a neutron into a proton, thereby optimizing
values of ĒB along an isobar parabola (5).
There is just one exception – technetium. No stable isotopes of this element of Z = 43 exist.
10.7 Example applications 303
The successive steps involved in follow-up processes of spontaneous fission are char-
acterized by individual velocities listed in Tab. 10.5 and illustrated in Fig. 10.18.
Fig. 10.18: Successive steps following initial fission events for binary fission. The division of a
fissile nuclide (1) (in parallel to α-emission (2)) creates a larger primary fragment (3) plus a smaller
primary fragment (not shown). This happens within ca. 10−14 s. These primary fragments de-excite
within a period of about 10−11 s and separate some prompt neutrons at high kinetic energy (4).
Only at this stage do primary nuclear products appear (5), which next continue to stabilize along an
isobar line by successive β−-transformation (6), thereby forming secondary products as
exemplified for mass number A = 131. The transformation along this isobar includes e.g. 131I, and
terminates at the stable nuclide 131Xe.
Tab. 10.5: Orders of magnitude of events leading to fission and following fission.
Process Time
Necking − ≈− s
be installed in a laboratory as a neutron source and can be used for various kinds
of research and training. 252Cf is also used as a primary neutron source in nuclear
reactors.
252Cf
2 645 a
2.645
6.118, 6.076, …
sf
, e- Fig. 10.19: Transformation properties of the spontaneous
fission-based neutron source 252Cf.
10.8 Outlook
Deformation of an unstable nuclide ✶K1 may yield a new nuclear ground state. For
this degree of deformation, higher energy states may exist, and the nucleus may
exist at an excited state.21 From this excited state, de-excitation may proceed “back-
wards” to the ground state, but also “forwards” to spontaneous fission as first de-
tected by S M POLIKANOV et al. in 1962.
Conceived of as “super-deformed” due to an extreme ratio of about 2:1 between the long and
short axes of the rotational ellipsoid.
10.8 Outlook 305
These excited states represent isomers of the same nuclide, and their process of
spontaneous fission is characterized by their own parameters of transformation.
Similar to the processes described above, they experience tunneling from the ex-
cited state through the fission barrier to the scission point, as per Fig. 10.20. Be-
cause of differences between the ground and excited states energies, as well as the
corresponding energies of the fission barriers, the half-lives of spontaneous fission
isomers are different from the half-lives of the ground states. For example, the half-
life of ground state 238U is 4.468⋅109 years, while the half-life of its spontaneous
fission isomer is 128 ns.22
1 3
potential energy U (MeV)
degree of deformation
Fig. 10.20: Concept of isomers in spontaneous fission. The excited state may remain for a given
period of time as a fission isomer. It proceeds further towards the scission point by tunneling
through the corresponding fission barrier (2), which is lower than the fission barrier for the ground
state (1). As this saves energy (3 and 4), spontaneous fission starting from a fission isomer state
proceeds faster compared to the ground state situation.
The existence of these fission isomers is indicated by white fields within the nuclide boxes of
the Chart of Nuclides, as per Fig. 10.4.
306 10 Spontaneous fission
Induced fission was first experimentally realized by German chemists HAHN (Nobel Prize in
Chemistry to HAHN in 1944 for “his discovery of the fission of heavy nuclei”) and STRASSMANN and
immediately accompanied by the theory of MEITNER and FRISCH.
However, several parameters can practically prevent this chain process: the kinetic energy of
the neutrons released, their number (flux or density in terms of neutrons per second and cm2) as
well as the number of 235U nuclides present within the range of the neutron initially emitted. There
is evidence that, due to specific local circumstances, such a process indeed happened in the ura-
nium mines of Oklo, Gabon (Africa) about 2⋅109 years ago.
11 Secondary transformations
https://doi.org/10.1515/9783110742725-011
308 11 Secondary transformations
introduce even more of new modes of emissions. The whole spectrum of these differ-
ent emissions thus must be considered in order to characterize a specific unstable
nuclide and to understand what is called “radioactivity”.
Unstable nuclei convert into stable ones by minimizing their absolute mass,
which is mirrored in terms of nucleon binding energy. Primary processes modify the
proton, neutron and/or overall mass numbers of the nuclides involved. Most of the pri-
mary processes are accompanied by the emission of small components x, i.e. beta par-
ticle, electron neutrino and electron antineutrino, alpha particle or neutron. Figure 11.1
illustrates the main directions of transformation of an unstable nuclide within the
Chart of Nuclides and their respective major emissions. In contrast, the defining fea-
ture of secondary transformations is that the values of Z, A and N do not change, as
per Table 6.3.
Fig. 11.1: Primary transformations of unstable nuclides indicating the three major emissions
relevant to naturally occurring nuclides. The illustration indicates the changes of nuclide positions
within the Chart of Nuclides due to changes in nucleon compositions.
11.1 From primary to secondary transformations 309
In many cases, a primary transformation of ✶K1 does not lead (and cannot lead, see
below) to the ground state of the new nuclide K2, eq. (11.1), shown in Fig. 6.22. In-
stead, individual excited states ⊙iK2 are populated1 as shown in eq. (11.2). Conse-
quently, the newly formed nucleus ⊙iK2 may exist in an “excited” state for shorter
or longer periods of time (see below), and the excited nucleons de-excite to levels of
lower energy according to the shell model of the nucleus.
Excited electrons of an atomic shell return to the ground state shell according
to the nominal electron shell profile of the atom in its corresponding chemical state
under emission of electromagnetic radiation (see Chapter 1). In a similar fashion,
the nucleons “move” towards lower-energy nuclear levels. The transitions may pro-
ceed from a higher-energy nuclear state to a lower-energy excited nuclear state (eq.
(11.3)) and/or from a given excited state to the final ground state shell occupancy of
the nucleus gK2 (eq. (11.4)). Secondary transformations proceed exothermically, and
similar to conventional chemical reactions, exothermic processes yield an excess of
energy ΔE. The corresponding emissions are denoted here as y. In most cases, y is
electromagnetic energy (or photons), typically referred to as “γ-radiation”.
✶
K1 ! g
K2 + x + ΔE (11:1)
✶ i
K1 ! K2 + x + ΔE (11:2)
i j
K2 ! K2 + y + ΔE (11:3)
i
K2 ! g K2 + y + ΔE (11:4)
Secondary transitions exclusively start from excited levels of nuclei – see also
Fig. 11.2. Suppose a primary transformation of ✶K1 leads directly to the ground state
g
K2 (eq. (11.1)), meaning no excited states of K2 are populated; in this case, no second-
ary process is needed at all. Thus, γ-emission in radioactive transformation processes
is not a singular process, i.e. does not belong to the group of primary β-particle- or
α-emissions.
In general, the larger the nucleus, the larger the number of possible excited states. These are
defined by nucleon shell occupancies different to the ground state, and, consequently, by individ-
ual quantum mechanics characteristics such as overall spin numbers, parity, but also by energy.
310 11 Secondary transformations
TRANSFORMATION TRANSFORMATION
TO GROUND STATE THROUGH EXCITED STATES
*K1 *K1
2K2
E covered by secondary
transition
1K2
K2 K2
gK2 // gK2
Fig. 11.2: Relation between primary and secondary transformations. Suppose a primary
transformation of ✶K1 leads directly to the ground state of gK2; then, there is no γ-emission at all
(left). Secondary transitions cover the transition from an excited level of ⊙iK2 to lower energy or
ground state levels of one and the same nuclide K2. The difference in energy ΔE between these
nuclear levels of K2 is typically emitted as γ-radiation, but also by other modes (right).
Altogether, there are three principal routes through which the difference in energy
ΔE between different excited levels (or one exited level and the ground state) of a
given nucleus can be expressed, as shown in Tab. 11.1.
Tab. 11.1: Three sorts of secondary processes transferring specific values of ΔE.
Fig. 11.3: Illustration of the successive character of primary and secondary transitions and
summary of the three options of secondary processes. Secondary processes can proceed via
emission of photons, release of internal conversion electrons and/or creation (formation) of an
electron + positron pair. Note that all three pathways exclusively originate from excited states of
⊙
K2 formed by primary nuclear transformations (or nuclear reaction processes). They proceed
within one and the same nuclide and may occur simultaneously.
The opposite effect will be discussed in Chapter 12, namely the metamorphosis of matter into
electromagnetic energy: matter (electron) and antimatter (positron) combine to 1.022 MeV energy.
This is called “annihilation”.
312 11 Secondary transformations
Similar to electron transitions in the shell of an atom, each nuclear transition has its
individual and discrete amount of energy, Fig. 11.4. The difference in energy released
in terms of electromagnetic radiation may be defined as ⊙iΔE = Qγ.
Fig. 11.4: De-excitation of an excited nuclear state to a ground state level via emission of a photon.
The photon is thus a package of energy, carrying away the differences in binding en-
ergy (here for nucleons in the nucleus, earlier for electrons in the shell) between one
discrete excited nuclear state to another, lower energy one. Figure 11.5 illustrates the
de-excitation of nuclear states of nucleons populated by primary transformations.
The primary transformation of unstable nuclei ✶K1 into more stable or really stable
nuclei, i.e. the change in nucleon composition, is a more or less “invisible” process.
In contrast, the simultaneous ejection of beta electrons or alpha particles during pri-
mary transformations,3 and now photons emitted during secondary transformation
processes, is directly measurable. This is summarized in the well-known picture of the
emissions α, β and γ, Fig. 11.6. The emissions can be detected quantitatively and used
to identify the kind of radionuclide and its absolute activity. However, γ-emissions4
are those most commonly used in nuclear and radiochemistry.
Of course, there are several more emissions already introduced, such as the positron and elec-
tron-(anti)neutrino (emitted in the course of β+- and electron capture transformations), clusters
larger than the α-particle, and protons and neutrons as ejected from extremely proton-rich or neu-
tron-rich artificially produced nuclides close to the corresponding nucleon drip lines.
Historically, these radiations are called α, β and γ, and have been identified in order of discovery
from naturally occurring unstable nuclides. Following the Greek alphabet, the first radiation char-
acterized was α-emission. The next generation of emission analyzed this way was named β. The
11.2 Photon emission 313
NUCLEON SHELLS
transition of nucleons
to lower-energetic levels of the nucleus
following a preceding primary nuclear transformation
4K2 4K2
E
3K2 3K2
E
2K2 2K2
E
1K2 1 K2
E
g K2
E g K2
Energy E of photons emitted
E = h·
E = E iK2 – E (i-j) K2
Fig. 11.5: Transition of nucleons (i.e. protons or neutrons) from excited nuclear states to lower
energy excited ones or to the ground state of the nucleus. Each nuclear state represents a
characteristic profile of nucleon shell occupancies. Following a primary nuclear transformation, a
nucleon may populate a higher energy (sub)shell and relax to a lower energy shell in a de-
excitation cascade, finally reaching the initial ground state. The individual differences in binding
energy are released as photons. This is comparable to the behaviour of excited electrons within
electron shells, as in the hydrogen emission spectrum.
final one categorized this way was the γ-emission, identified in 1900 by P VILLARD when measuring
the different types of radiation emitted from radium. The term “γ-rays” was introduced by E RUTH-
ERFORD in 1903 (in fact, RUTHERFORD named them all: α, β and γ emissions). Photons are symbolized
by the Greek letter γ in nuclear sciences. In other sciences such as chemistry, the expression hν is
used.
φῶς (Greek) = phôs = light.
314 11 Secondary transformations
Fig. 11.6: Most relevant “radioactive” emissions (or “radiative” emissions) in nuclear and
radiochemistry. They are schematically categorized in terms of deviation in an electromagnetic
field, reflecting the charge and mass of each emission. For γ-emission, there is no charge, no mass
and thus no deviation. In contrast, α-particles are positively charged and have a relatively heavy
mass, while β−-electrons are negatively charged and about 4000× less massive than the α-particle.
The photon is the quantum of electromagnetic radiation; this is true for all spectral
regions, including visible, ultraviolet and infrared light, radio waves, microwaves,
and those waves originating from nuclear transformations. Photons travel with the
maximum velocity of the universe, which is one of the fundamental constants in
science: the speed of light c = 299,792,458 m/s, which is 1.079⋅109 km/h. The energy
of a photon can be related to its frequency ν, wavelength λ (ν = 1/λ), the PLANCK con-
stant h, angular frequency ω, the reduced PLANCK constant ħ = h/2π and the speed c
of photons (ν = c/λ) via eq. (11.5).
c
E = hν = h = hω (11:5)
λ
The spectrum of electromagnetic radiation is huge, as shown in Fig. 11.7. In terms
of wavelength (λ in m) or frequency (ν in s−1, which is Hz) and in terms of energy
(in eV), it covers many orders of magnitude: from λ ca. 10 pm to 1 km, from ν ca.
1020 to 105 Hz, and E from keV to MeV. Only a limited part of the spectrum is visible
11.2 Photon emission 315
to the human eye: about 400 to 700 nm in wavelength. This is neighbored by the
ultraviolet and infrared regions, i.e. down to about 10 nm and up to about 10 μm,
respectively. Beyond and towards longer wavelengths, there is, for example, the do-
main of radio waves covering an area of about 1000 MHz (UHF) to about 50 MHz
(VHF). In the context of radioactive secondary transitions and post-processes, the
high frequency and energy (i.e., short wavelength) branch of the electromagnetic
spectrum becomes relevant. It is divided into two sub-parts: γ-radiation and X-rays.
A good reason to make a distinction here is not the value of frequency, wavelength
and energy, but rather their origin.
11.2.3.1 γ-rays
Gamma radiation6 is understood to originate from the nucleus of a nuclide via de-
excitation of an excited level of the nucleus. In terms of wavelength, it is approximately
<10−11 m (which is <10 pm or <0.1 Å, i.e. shorter than the diameter of a nucleus); in terms
of energy, it is >0.1 MeV, as summarized in Tab. 11.2.
11.2.3.2 X-rays
In the framework of radioactive processes, electromagnetic radiation emitted from
the shell of the atom (i.e., not from the nucleus) is signified as X-radiation.7 Com-
pared to γ-rays, the wavelengths are longer (ca. 0.01–10 nm), and energies are lower
(ca. 0.1 keV to ca. 0.1 MeV). Thus, the domain of X-rays is between that of γ-rays and
γ-Rays X-rays
ultra-violet
visible ligtht
infra-red
10 -14 10 -12 10 -10 10 -8 10 -6 10-4 10 -2 10 0 10 2 10 4
Wavelength ( m)
X -ray
-ray
UV light. In terms of energy, X-rays are also discriminated by their penetration power
into “soft” X-rays (energies < 10 keV) and “hard” X-rays (of about 10 to 120 keV).
An ejected γ-photon contains almost all the energy released in a transition between
two nuclear states. Similar to the small fragment x released in a primary transfor-
mation (x = α- and β-particles), the nucleus also experiences a relatively low recoil
energy during secondary processes. The overall energy available within the transi-
tion is Qγ in the present case, and is between ca. 0.1 and 10 MeV. The value for
RECOIL
EK2 is derived from the conservation of the two momenta as pγ = precoil. With
E = hν and pγ2 = hν/c, eq. (11.6) gives a relationship between the mass number A of
the recoil nucleus and the energy of the photon emitted:
The question is: do these six options proceed with identical probabilities, or are some
transition steps preferred over others? If the latter is true, then why? Indeed, all the
potential steps have an individual branching. The reason for this defined protocol
takes us back to the nuclear structure of the individual excited states, back to the
shell model of the nucleus, i.e. back to the intrinsic quantum physical parameters of
the nucleons and the resulting overall parameters of the nuclear state: overall orbital
momentum (spin) and parity. Similar to β- and α-transformations, the process in sec-
ondary photon transitions is also defined by initial and final wave functions and a
transition matrix element. The quantum physical parameters needed for each initial
and final state and for the differences ΔE between the two states are:
– the orbital quantum number l,
– the magnetic orbital quantum number,
– the spin quantum number s and the resulting overall orbital momentum L,
– the overall angular momentum S,
– the overall momentum J obtained from orbital-spin coupling.
The higher the mass number of an unstable nucleus (which reflects high Z and N numbers), the
higher the number of potential excited levels, and the more possible secondary transition steps.
318 11 Secondary transformations
K2initial Ji , i
| Ji – Jf | ħ = J
K2final Jf , f
Fig. 11.8: De-excitation between two excited nuclear states defined by quantum physical
parameters J and Π for the initial and final states. The transition via photon emission proceeds for
specific values of ΔJ, and the photon must carry away this difference in overall momentum.
Nuclei with overall orbital momentum 0 show a monopole electromagnetic moment, and be-
cause the angular momentum of a photon is 1, these transitions are impossible.
This defines a gyromagnetic parameter, which is the ratio (or proportionality factor) between a
magnetic dipole moment and an angular momentum of a system (given by units of C / kg).
11.2 Photon emission 319
11.2.5.4 Parity Π
The parity of the photon is negative, while the parities of the electron, neutron and
proton are positive. For each secondary transition, the parity of the total system
must be conserved. However, changes in overall orbital momentum between initial
and final states act in an alternating way for electric and magnetic multipole radia-
tion, according to the photon’s parity as either (−1)l or (−1)l + 1, respectively. This is
shown in eqs. (11.8) and (11.9).
El Πi = ð − 1Þl Πf (11:8)
Ml Πi = ð − 1Þl + 1 Πf (11:9)
Thus, for the same order of the photon’s multipole, the parity of electric multipole ra-
diation is opposite to the parity of magnetic multipole radiation. Consequently, multi-
pole radiations of type E1, M2, E3, etc. stand for transitions changing the parity, while
in the case of M1, E2, M3, etc. the parity is not changed between initial and final nu-
clear states. Combinations of ΔΠ = “no” and ΔJ = 2, 4, etc. determine electric MP radia-
tion, while combinations of ΔΠ = “no” and ΔJ = 1, 3, etc. determine magnetic MP
radiation, shown in Tab. 11.3. If changes in ΔJ between initial and final nuclear states
are integer or noninteger, there are two options depending on whether the overall par-
ity remains (ΔΠ = “no”) or changes (ΔΠ = “yes”). Values of ΔJ define dipole and quad-
rupole etc. monopoles of the corresponding type. For example, ΔJ = 1 gives either E1
or M1, depending on whether the parity is changed (ΔΠ = “yes”) or not (ΔΠ = “no”).
Figure 11.9 illustrates three examples of photon transitions.
Tab. 11.3: Types of photon multipole (MP) radiation in secondary transitions. IΔJI
indicates the absolute value of the difference.
El IΔJI ΔΠ Ml IΔJI ΔΠ
di- E yes M no
quadru- E no M yes
octu- E yes M no
hexadeca- E no M yes
1+ 2+ ½-
= yes
= no
= no
l=1 E1
l=1 M1 l=2 E2
l=2 M2
2 ≤ l ≤ 2;
1 ≤ l ≤ 1;
1 ≤ l ≤ 2;
0- 0+ 3/ 2 -
A B C
Fig. 11.9: Examples of photon transitions with different combinations of ΔJ and ΔΠ. For (A) l = 1,
ΔΠ = yes and for (B) are l = 2, ΔΠ = no. For each of (A) and (B), there is only one option, namely M1
and E2, respectively. For the case (C) with l = 1 and 2, and with ΔΠ = yes, there are two transitions,
namely E1 and M2; yet with a very different probability in favor of E1.11
radiation) and ΔΠ = “no”. Yet, all of the transitions are allowed to occur, though with a
very different branching. With increasing order 2l, the probability of these photon emis-
sions decreases successively.
Example (A) could represent the transition of an excited level of 26Mg (formed by primary pro-
cesses of β+- and ε-transformation from excited 26Al, t½ = 7.2⋅105 years) to its ground state with
t½ = 0.49 ps. It emits a photon of Eγ = 1.809 MeV, which is a sensitive measure of the astrophysi-
cal processes occurring in supernovae. Example (B) shows transitions which occur for instance
for 58Co ! β− ! 58Fe. The two excited states are ⊙2(58Fe) = 2+ and ⊙1(58Fe) = 2+, while the ground
state is of 0+. The primary transformation populates ⊙1(58Fe) with 99.48% efficiency. The main
MP radiation is E2. Example (C) is for two options in ΔJ, namely 1 and 2, and therefore two differ-
ent transition options arise.
11.2 Photon emission 321
ΔJ = , , . . . ΔJ = , , . . . Πi = Πf “no” or “ −”
ΔJ = , , . . . ΔJ = , , . . . Πi ≠ Πf “yes” or “ +”
2.4Sr2
λel.MP = 1021 s − 1 (11:10)
ðE=197Þ2l + 1
0.55SA − 2=3 r2
λmagn.MP = 1021 s − 1 (11:11)
ðE=197Þ2l + 1
2
2ðl + 1Þ 3
S= (11:12)
lf135 . . . ð2l + 1Þg2 l+3
Values of S are thus only a function of l. For l = 1, 2, 3 and 4 the corresponding values are
2.5⋅10−1, 4.8⋅10−3, 6.25⋅10−5 and 5.3⋅10−7, respectively.
. . . which is similar to those of primary transformations; see eq. SARGENT-type and GEIGER/NUT-
TAL-type correlations.
322 11 Secondary transformations
Thus, the fastest transitions at high Eγ are those for E1 (ca. 10−16 s); conversely,
those for M3 are very slow (ca. 10−4 s). For higher MP, transition velocities are not of
seconds, but years and even thousands of years, i.e. proceed only with extremely
low probability.
11.2.6 γ-spectra
Individual transitions between discrete excited states or from one excited state to the
ground state show γ-emissions of individual and discrete energy. In a γ-spectrum,
these are recorded according to their energy and relative abundance. Figure 11.10 il-
lustrates the concept for some select transitions. Whereas the spectrum arranges the
transitions according to increasing γ-energy (i.e. iΔE values), their intensities may be
very different.
Let us discuss the secondary transformations following the primary β−-
transformation of 131I. The most frequently occurring primary transformation (89.4%)
is the one populating the excited state ⊙4 (among 8 different nuclear levels of
131
Xe) – see Table 8.7. The de-excitations originating from this excited state will pre-
dominantly determine the γ-spectrum. The question is, which lower levels are ad-
dressed most often? The corresponding values of JΠ for levels 4, 3, 2, 1 and the
ground state are 5/2+, 9/2−, 11/2−, ½+ and 3/2+, respectively (see Table 8.7).
11.2 Photon emission 323
Fig. 11.10: The origins of a γ-spectrum. Transitions between discrete excited states or to the ground
state show the hypothetical γ-emissions of individual discrete energy (left). In a γ-spectrum (right),
the photons are recorded according to their energy (x-axis) and relative abundance or intensity
(y-axis). In the artificial example given, three photon emissions are selected. The most energetic is
(1) with ΔE = ⊙4EK2 − gEK2, but it proceeds at a relatively low probability. Transition (2) has a lower
value of ΔE = ⊙2EK2 − gEK2, thus the photon emitted is of lower energy. It is, however, the most
probable transition of the three. Transition (3) stands for ΔE = ⊙2EK2 − ⊙1EK2, thus the photon
emitted is of lowest energy. This is obvious from the point of view of energy. But what about the
intensity? There are two factors to consider, namely: (i) how populated the initial excited state was,
and (ii) what its probability of relaxing to a certain lower-energy nuclear state was, according to
selection rules. Note that later, the efficacy of the radiation detector will also become a factor, as it
is not identical for all photon energies.
The most relevant ones are M1 + E2 for γ[⊙4,g] (making up 83.1% of the photon emis-
sions from all excited levels) and E2 for γ[⊙4,⊙1] (making up 6.36% of the photon
emissions from all excited levels). Subsequent to the photon emission γ[⊙4,⊙1],
324 11 Secondary transformations
ΔE (keV)
K2 = Eγ (keV)
JΠ relative to t½
¹³¹Xe and photon multipole
ground state
100.0 636.989 E2
0.005 85.9
0.9 358.4
0.044 232.18
1.32 302.4
0.025 295.8
722.909
⁸K2 5/2+ 0.53 ps
7.51 284.305 E2
404.815
⁵K2 3/2+ 75 ps
100 163.930 M4
346.490
⁴K2 5/2+ 67.5 ps
163.931
²K2 11/2– 11.93 d
(meta-stable)
80.185
¹K2 1/2+ 454 ps
3
ᵍK2 3/2+ 0 stable
Fig. 11.11: Individual secondary photon transitions between discrete excited states of 131Xe as
populated from β−-transformation of 131I. Excited nuclear states and electric and magnetic
multipoles are given for each transition, together with the energy of the corresponding photon. The
most relevant excited nuclear state is ⊙4(131Xe) populated by 89.4% of all β−-transformations (red
line). From each excited nuclear state, secondary transitions through emission of photons proceed
simultaneously, though with specific half-lives, indicated in the right-hand column. Starting from
⊙4 131
( Xe), the photon emissions would be γ[⊙4,⊙3], (5/2+ ! 9/2−; 2 ≤ ΔJ ≤ 7; ΔΠ = yes),
γ[⊙4,⊙2] (5/2+ ! 11/2−; 3 ≤ ΔJ ≤ 8; ΔΠ = yes), γ[⊙4,⊙1] (5/2+ ! ½+; 2 ≤ ΔJ ≤ 3; ΔΠ = no) and γ[⊙4,g]
(5/2+ ! 3/2+ ground state; 1 ≤ ΔJ ≤ 4; ΔΠ = no), respectively. Here, γ[⊙4,g] and γ[⊙4,⊙1] (wavy red
lines) are relevant because of relatively low changes in overall momentum and conservation of
parity. From RB Firestone, VS Shiley (eds.), Table of Isotopes, 8th edition, John Wiley & Sons,
New York/Chichester/Brisbane/Toronto/Singapore, 1996.
there is the one of γ[⊙1,g] of type M1 + M2 and with 6.72% abundance. The individual
half-lives of excited levels 4 and 1, according to eqs. (11.10) and (11.11), are 67.5 ps
and 454 ps.
The corresponding energies of the different photons are 364.49 keV and 284.31 keV,
respectively. Individual probabilities are relative to the dominant transition of γ[⊙4,g]
(=100%) and γ[⊙4,⊙1] (= 7.51%), respectively. Referring back to the 89.4% initial popula-
tion of the ⊙4(131Xe) excited level via primary β−-transformation of 131I, this gives overall
values of photon emissions for γ[⊙4,g] = 83.1% and γ[⊙4,⊙1] = 6.36%, respectively. Usu-
ally, these individual percentages are given in terms of photon emission coefficients
11.2 Photon emission 325
εemission, which in this case are εγ = 0.831 and εγ = 0.0636.14 This finally is reflected in
the γ-spectrum of 131I, Fig. 11.12. Note that the count rates measured here originate not
from 131I directly, but from individual excited nuclear states of 131Xe.
32768
722.1 keV
502.7 keV
284.3 keV
80.3 keV
636.5 keV
364.5 keV
32768
8192
512
24756 128
32
d
ounts per second
8
364.5 keV
131Xe
57.4 d
L (35)
e
722.1 keV
284.3 keV
80.3 keV
636.5 keV
502.7 keV
8192
k
131I
k
8.02 d
0.6, 0.8 ...
0 200 400 600 800 1000 1200 1400 1600 1800 2000
-energy (keV)
Fig. 11.12: The γ-spectrum of 131I. Note the linear scale and the logarithmic scale (insert) of the
y-axis. The two most abundant photon emissions are 364.489 keV and 284.305 keV. According to
Fig. 11.11, other γ-emission energies are less probable, yet well identified when visualized via the
semilog plot.
. . . and are needed to transfer experimentally measured count rates into absolute radioactivity
of a nuclide sample; see eq. (5.11) and Chapter 1 in Volume II.
326 11 Secondary transformations
Emission of discrete photons from excited nuclear levels is by far the most frequently
occurring pathway of secondary transitions. But there are also two more options. The
first is called “inner conversion” or “internal conversion” (IC), Fig. 11.13.
e-
IC
Fig. 11.13: Internal conversion. Electromagnetic energy of ⊙iΔE is not emitted as a photon, but
converted to a shell atomic electron, which is subsequently released from its shell and leaves the
atom.
It represents a situation in which the difference in energy ⊙iΔE between two nuclear
levels is transferred to an atomic shell electron.15 Similar to the electron capture pro-
cess in primary β-transformations, this can only be understood in the context of
quantum physics: s-electron orbital wave functions show a certain probability of ex-
isting close to and even inside the nucleus of an atom. Here it directly gains the
equivalent of ⊙iΔE, provided this amount of energy is larger than the electron bind-
ing energy of that electron shell EB(e). Consequently, all internal conversion electrons
are former s-orbital electrons, and most of these originate from an s-orbital as “close”
to the nucleus as possible. The kinetic energy EIC of the ejected electron is equivalent
Internal conversion is well known in conventional chemistry, when high-energy states of a mol-
ecule or an atom may transit to lower energy states with or without emitting photons. In the latter
case, this energy is absorbed internally, e.g. affecting vibrational systems.
11.3 Conversion electrons 327
to ⊙iΔE minus the energy needed to overcome the electron binding energy, as per
Fig. 11.14 and eq. (11.13). Internal conversion electrons thus have their own discrete
energy – different from β−-particles.
i
EIC = ΔE − EBðeÞ (11:13)
The IC electron is not excited to a higher energy electron shell – it is released from
the atom.16 Overall, no photon is emitted, because one was never created.17 Internal
conversion refers to the origin of the energy that is created inside the nucleus,
rather than deposited from external sources (such as in Roentgen spectroscopy).
K
e-
L
M
e-
e-
Fig. 11.14: Origin of internal conversion electrons from individual s-shell orbitals of individual main
quantum number n, i.e. K-shell, L-shell etc.
Note that this is the emission of a “real” atomic electron, i.e. one which already existed in a
lower electron shell of the atom. This is in contrast to the emission of electrons as β-particles, cre-
ated instantaneously in the nucleus of that atom over the course of the β-transformation.
Note that this internal conversion should not be confused with the “photoelectric effect”. In the
latter case, an existing photon “hits” a shell electron and ejects it.
328 11 Secondary transformations
Kα Kα
IC(E1) IC(M1)
10³
10²
10¹
Internal conversion coefficients αIC
10¯¹
Z
10¯²
100
Z
10¯³ 100
10¯⁴ 40
40
20
10¯⁵ 20
10
10
10¯⁶
10⁰ 10¹ 10² 10³ 10⁰ 10¹ 10² 10³
(a) Eγ (keV) Eγ (keV)
Fig. 11.15: Internal conversion coefficients αIC depending on parameters of an excited nuclide: Z, Eγ,
multipolarity and shell number. ICC values are higher for low γ-energies, for electric
(vs. magnetic) multipoles, for higher multipolarities and for KαIC relative to LαIC. From: IM Band, MB
Trzhhaskovskaya, CW Nestor, PO Tikkanen, S Raman, Dirac–Fock Internal Conversion Coefficients,
Atomic Data and Nuclear Data Tables 81 (2002) 1–334.
11.3 Conversion electrons 329
10³
LαIC(E1) KαIC(E3)
10²
10⁶
10¹
10⁴
10⁰
Internal conversion coefficients αIC
10¯²
10⁰
10¯³
Z
100
10¯⁴
Z
10¯²
100
10¯⁵
40
40
10¯⁴
10¯⁶ 20
20
10¯⁷ 10
10
10¯⁶
10⁰ 10¹ 10² 10³ 10⁰ 10¹ 10² 10³
(b) Eγ (keV) Eγ (keV)
Internal conversion is a domain for low energy ⊙iΔE transitions, particularly for
monopole modes of 0+ ! 0+ where γ-emission is impossible. Nevertheless, internal
conversion is also possible for multipole transitions.
330 11 Secondary transformations
IC is most often observed for 1s-K electrons, but also with a lower probability for L, M
and N-shells. In all cases except for ΔJ = 0, internal conversion accompanies photon
emission from excited nuclear states. The two subtypes combine their individual proba-
bilities. Internal conversion coefficients (ICC) αIC are derived by eq. (11.14), defined as
the ratio of the number of electrons NIC ejected from a certain atomic shell to the number
of photons Nγ simultaneously leaving the atom. According to the density of wave func-
tions of K- and L-shell electrons, IC dominates for K-shell at a ratio of KαIC > LαIC (ca. 8:1).
NIC
αIC = (11:14)
Nγ
Moreover, the density of wave functions of the same atomic electron shell is strongly
influenced by the proton number of the nucleus, which thus has an impact on the
internal conversion coefficients. In general, IC dominates for increasing Z (because of
the higher probability for s-orbital electrons to exist close to/within the nucleus) and
relatively low values of ⊙iΔE (0.2 MeV and lower, according to KαIC ≈ Z3/Eγ). The mul-
tipole order and photon energies also influence these coefficients. Figure 11.15 com-
pares values of αIC calculated for different Z, Eγ and shell number values for E1 and M1
multipolarity transitions.
ected
K
ectrons dete
CONVERSION ELECTRONS
L
umber of ele
- SPECTRUM M
relative nu
E max
In order to conserve energy, the value of ⊙iΔE must at least equal the mass of
the two electrons. Electron masses are 0.511 MeV (or 0.00055 u) each, so the overall
energy follows ⊙iΔE ≥ 1.022 MeV.
Pair formation thus exclusively emerges at ⊙iΔE > 1.022 MeV, i.e. for a relatively
high difference between the two nuclear energy levels involved. It is thus a rela-
tively rare effect. Internal conversion, in contrast, arises at low values of ⊙iΔE, and
a large number of radionuclides emit IC electrons. The emission of a γ-photon is,
however, most frequent. It takes place over the whole range of iΔE values, and rep-
resents several thousand cases. Figure 11.18 illustrates the range of the three sec-
ondary transition options depending on the value of ⊙iΔE.
e- e+
IC
-emission
pair formation
Fig. 11.18: Appearances of the three secondary transitions depending on the values of iΔE.
excitation between individual excited states or the final de-excitation from one excited
state to the ground state is very fast, typically lasting only 10−16 to 10−13 s. The overall
secondary transformation is thus extremely fast, even when cascades of several transi-
tions are involved.
Suppose for the reasons introduced in Figs. 6.22 and 6.23, one (or more) of the indi-
vidual excited nuclear levels of K2 “remain” for longer periods of time. Under these
circumstances, this excited state ⊙iK2 will have a much longer half-life than the other
excited levels of ⊙K2 – about five orders of magnitude longer than the next longest-
lived transition.18 These half-lives represent a particular state in a cascade of second-
ary transitions where time seems to stand still for a while (see Fig. 11.19).
This happens in the context of the selection rules, i.e. whenever the differences in
overall angular momentum are large (octa-, hexadeca-, or higher multipole orders) or
when parity is violated. The phenomenon is often observed for very small values of
⊙i
ΔE between the initial and final states. This excited state is not really stable like the
final, ground state of gK2, but remains intact for a significant period of time. Because
it lies between unstable and stable (ground state) levels, it is called “metastable”,
m
K2. Compared to the ground state, it reflects a nucleus of identical nucleon composi-
tion in terms of A, Z, and N, and is therefore also referred to as “nuclear isomer”.18
There are many metastable isomers with half-lives in the range of minutes, hours,
days and years.19
Figure 11.11 showed the individual half-lives of the various excited levels in
⊙ 131
( Xe). One of them, ⊙2(131Xe), was of t½ = 12.93 days, i.e. much longer than the
other (extremely fast) transitions with nano- and picosecond half-lives. Indeed, the
second excited level of 131Xe represents such an isomer. Figure 11.20 shows the two
pairs within the primary transformations of 131I ! β− ! 131Xe and 81gRb ! ε,β+ !
81m
Kr/81gKr. There are about 700 of these metastable isomers. Some are of interest in
fundamental research; others are relevant to important applications.
For nuclides of Z > 92, a metastable nuclear isomer may transform via spontane-
ous fission, then called a “fission isomer”.20 The first sf isomers appear as 236mU
(t½ = 120 ns) and 238mU (t½ = 298 ns). 242Am has a fission isomer 242mAm with a
long half-life of 141 a. Interestingly, this half-life is longer lived than the ground
state of 242gAm of t½ = 16 h.
Nuclear isomers are thus just another version of isomerism, in addition to the various isomer
types known in classical chemistry. Stereoisomers, for example, are identical in the number of,
identity of and chemical bonding between atoms, but are different in the orientation of those atoms
in space. Organic compounds that contain a chiral carbon usually have two non-superimposable
structures forming enantiomers. Just like our left and right hands, these two structures are mirror
images of each other and are otherwise structurally the same – but may differ in some chemical
properties. For example, carvone forms two enantiomers, namely R-(–)-carvone and S-(+)-carvone.
The first one smells like spearmint, the other like caraway.
The ground state 180gTa has a half-life of 8.125 h. One excited nuclear state 180mTa remains at its
excited level with t½ > 1015 years because of the high ΔJ = 8.
First discovered for 234mPa (t½ = 1.17 m) by OTTO HAHN in 1921 when studying the β−-processes
of 234gPa (t½ = 6.70 h), the second product of the naturally occurring 238U chain.
334 11 Secondary transformations
*K1
*iK2
m K2
t½ ≈ 10-9 s –1015 a
E
g K2
time
Fig. 11.19: Metastable excited nuclear states mK2. One (or more) individual levels ⊙iK2 shows a
significantly prolonged half-life relative to the longest-lived of the other, very short-lived secondary
transitions. The simplified representation assumes that velocities of a transforming nucleus follow
the idea of a ball bouncing downstairs. Whenever the width of a step is small and the next stair is
much deeper, the ball moves very fast (like the hillside model of the valley of β-stability concept,
as in Fig. 6.13). If the ball meets a step which is only a bit deeper than the initial one, and much
broader, its speed decreases.
Among the three subtypes of secondary transitions, photon emission is the most rele-
vant mode. The photons emitted are relatively easy to register quantitatively by ade-
quate detectors (see Chapter 1 of Volume II). Compared to the emissions of the
11.6 Example applications of photon emission 335
131Xe
11.9 d
81Rb l 127
e-
13.1 s 2.3.105 a
l 190 ,
Fig. 11.20: Description of metastable isomers for secondary transitions in 131I ! β− ! 131Xe and
81g
Rb ! ε,β+ ! 81mKr/81gKr. For the primary transformations, the most relevant data are in the
boxes of the initial nuclide ✶K1, such as half-life and main β−-energies, which also includes the
main γ-energies of the subsequent secondary electromagnetic transitions. Within the Karlsruhe
Chart of Nuclides the isomers are small white fields within the boxes of the stable product
nuclides. They show the half-life of the nuclear isomer and parameters of transformation including
the γ-energy (in keV). 131I/ 131Xe represents a secondary process subsequent to a β−-primary
process.21
primary transformations, which (except electron neutrinos) all are particles, i.e.
α-particles, β-particles and neutrons, the photons interact with matter only to a very
small degree. They penetrate glassware and metallic containers and are only gradu-
ally affected in intensity, but not in energy. Consequently, photons are a noninvasive
measure of the identity of the radionuclide and represent the most valuable diagnos-
tic tool in nuclear research, nuclear technology and nuclear medicine. In the follow-
ing, some directions of applications of γ-emitting radionuclides are indicated. Most of
them will be discussed in more detail in Volume II.
In addition, photons can be used to noninvasively determine the absolute ra-
dioactivity of a sample within glass or metal vessels or containers. Photons emitted
in secondary transitions of unstable nuclides are thus the most accurate tool to
quantify nuclear transformation processes.
The 81gRb/ 81mKr process starts from the ground state of unstable 81gRb and populates the meta-
stable isomer 81mKr. Due to the half-lives of 4.58 h relative to 13.1 s, this is a radionuclide generator
system, used for diagnostic imaging in nuclear medicine.
336 11 Secondary transformations
*K1
K1
transsformation
E ccovered by
primary
p
iK2
ent secondaryy
E ccovered by
ansition
mK2
subseque
tra
, , sf
K3
Fig. 11.21: Options for metastable isomers. Secondary transformation to lower energy nuclear
states of ⊙K2 via e.g. photon emission vs. primary transformation to nuclide K3 (through α- or
β-transformation or by spontaneous fission).
11.6.1 Astrophysics
High-energy photons traverse the universe and the atmosphere of the earth, making
them useful in measuring astrophysical processes. 26Al for example has a half-life
of 7.16⋅105 years. When formed in supernovae, it transforms to stable 24Mg via an
excited nuclear level of 24Mg of 6.35 s half-life. From that excited nuclear state, the
subsequent secondary transition into the ground state 24gMg emits a photon of
1.809 MeV energy. Satellites register this γ-ray, emitted from the center of our
11.6 Example applications of photon emission 337
99Ru
99Tc
T
2 1 3
6.0 h 2.1.105 a
l 141 0.3
99Mo
66.0 h
1.2
740, 182, …
Fig. 11.22: Origin and fate of the nuclear isomer 99mTc. The isomer 99mTc is a small and basically
white box within the blue box (β−-transformation) of the unstable 99Tc. 99mTc transits to the ground
state 99gTc via secondary processes (1), mainly through emission of a 141 keV photon. The quantum
physical parameters belong to the γ[⊙1,g] transition. The main photon multipole is M1 + 3.3% of E2
with 1/2– and 9/2+ for 99mTc and 99gTc. Simultaneously, 99mTc transforms via β−-emission (blue
triangle in the white field) to the more stable nuclide of the A = 99 isobar, which is stable 99Ru (2).
99
Ru is thus predominantly obtained through the direct, primary β−-transformation of 99gTc (3), but
additionally, even if to a very small percentage, by primary β−-transformation from 99mTc.
galaxy, as shown in Fig. 11.23. These data can contribute to the understanding of
rotational movements of the galaxy.22
For further applications of radionuclides in the context of dating, see Volume II, Chapter 5.
338 11 Secondary transformations
ANTICENTER REGION
AND TAURUS CLOUDS
INNER GALXY CARINA VELA REGION
CGNUS REGION
Fig. 11.23: Map of the distribution of 26Al in the galaxy, indicating where in the Milky Way most
supernovae exploded during the last million years. Courtesy: MPE Garching, R Diehl, Comptel
1991–2000, ME 7, Plüschke et al. 2001.
11.6.3 Industry
The unique γ-energies, intensities and half-lives of photons serve as a fingerprint of the
isotope they were emitted from. They thus represent a valuable tool for quantitative
Nobel Prize in Physics, 1961, to RUDOLF MÖSSBAUER for “ . . . his researches concerning the reso-
nance absorption of gamma radiation and his discovery in this connection of the effect which bears
his name”.
11.6 Example applications of photon emission 339
For photons created in the course of secondary transformations, the medical technol-
ogies are scintigraphy and single photon emission computed tomography (SPECT).
Requirements are thus:
– a physical half-life adequate to the medical purpose,
– a photon energy in the range of ca. 100 to 400 keV adequate to the optimum
detection efficacy of the scintillation detectors used,
– the absence of further nuclear transitions (i.e. transformations should ideally
lead to a stable transformation product) and
– robust availability of the nuclide and its labeled compounds.
The most common examples are listed in Tab. 11.7. Clearly, the metastable, genera-
tor-derived 99mTc is the most important radioisotope in nuclear medicine and
SPECT imaging (Fig. 11.22). The interesting point here (different from the processes
illustrated for 131I, discussed below) is that this particular excited nuclear level has
an extended period of existence, cf. Fig. 11.19. It is a metastable nuclide of 99Tc and
is therefore denoted by 99mTc. Its half-life is 6.0 h. This metastable 99mTc undergoes
a secondary transition process to 99mTc, accompanied by the intense emission of
low energy photons of 140.5 keV (see Chapter 11). Thus, 99mTc is available from a
convenient radionuclide generator system, provides a daughter nuclide every day
(according to the mathematics of the transient generator equilibrium, see chapter
340 11 Secondary transformations
5.4.3.4), and represents a source of photon radiation very useful for medical diag-
nostic imaging via SPECT (single photon emission computed tomography). Today,
99m
Tc is the number one medical isotope, with more than ten thousand applications
for patient diagnoses worldwide every day.24
131
I is also applied in “diagnostic” activities, and is another key radioisotope in
nuclear medicine. It is often used on an industrial scale as a “medical” radionuclide
in the treatment of malignant tumors, or to treat diseases of the thyroid and other
glands. As the excited states of 131Xe de-excite, they emit electromagnetic radiation.
According to the dominant population of one state, the photons emitted from this
level approaching the ground state of 131Xe either directly or via another excited state
create intense photon emissions at 364 keV and 284 keV energy, respectively, as
shown in Figs. 11.11 and 11.12.25
Tumor treatment from external radiation source: In contrast to the low energy
photons of 99mTc and 131I, which for diagnostic purposes pass through the human
body without causing critical radiation doses, intense and high energy photons (Eγ >
1 MeV) of photon-emitting radionuclides are applied in cancer treatment by means of
external irradiation (“γ-knife”). Two examples are 60Co (t½ = 5.272 a, Eγ = 1332 and
1173 keV, shown in Fig. 8.21) and 137Cs. Clinically, enormous activities (several thou-
sand GBq) of 60Co are shielded in a medical device, designed to lead the photons in
sophisticated arrangements to well-localized brain tumors for noninvasive therapy.
Tumor treatment from internal radiation source: In contrast to the high energy
photons of e.g. 60Co, very low energy photons can be clinically generated from a
sealed radiation source, which is deposited in the tissue of interest. It is referred to as
“brachytherapy”, from the Greek word for “soft”. The most relevant radioisotopes to
this type of therapy are 131Cs (t½ = 9.7 d), 103Pd (t½ = 17.0 d), 125I (t½ = 59.6 d), and
192
Ir (t½ = 73.8 d).
For 99mTc to realize its full therapeutic potential still requires, in addition to the nuclear data,
the contribution of radiopharmaceutical chemistry. The challenge is to turn this radionuclide,
which does not exist naturally on earth, into biologically relevant molecules, thereby allowing for
the measurement of a large variety of physiological processes in the living human body.
Due to its availability from a 99Mo/99mTc radionuclide generator and the success of radiopharma-
ceutical chemistry to develop various biologically relevant 99mTc-labeled molecules, 99mTc-SPECT im-
aging and scintigraphy represent about 90% of all diagnostic studies in nuclear medicine and
account for over 80% of the overall worldwide use of radionuclides in medicine. These procedures
are performed on more than 300,000 patients worldwide every day (see Chapter 13 of Volume II.)
11.7 Outlook 341
11.7 Outlook
Photon emission and pair formation finally de-excite the intermediate of a primary
transformation to its stable ground state. In doing so, these two secondary transitions
terminate the nuclear transformation processes initiated by the primary transformation.
The unstable nuclide ✶K1 has been transformed into a new nuclide K2, either stable or
at its ground state of an unstable nuclide.
However, this is not the case for IC. Why? Internal conversion is the ejection of
an s-orbital electron, but leaves a vacancy (hole) in the electron shell of K2. The re-
maining atom is thus not yet “intact”. The same is true for one subtype of the primary
β-transformation, namely electron capture, which also created a hole in an atom’s
inner s-orbital electron shell. In both cases, the atom must now handle this situation
in another subsequent step, somehow “filling” the vacancy in its own electron shell!
What happens here is not part of a primary or secondary transformation, but a post-
effect; it is definitely part of the whole story of a radionuclide transformation.
Furthermore – there is a second group of post-effects. Recall that primary β+-
transformations create a positron. While the nucleus has achieved its new nucleon
configuration, the destiny of the positron still remains part of the overall transfor-
mation process. It is ready to annihilate with an electron, orbiting the shell of some
far-away atom. Consequently, annihilation radiation is added to the other emis-
sions characteristic for the primary transformation that created the positron, and
the secondary emissions as consequences of the primary one.
Similarly, electron + positron pair formation was one of the outcomes of a second-
ary pathway within the ⊙iK2 ! ⊙K2 or gK2 nuclear transformation. The pair of elec-
tron + positron is released from the nucleus, and immediately reacts further, inducing
a characteristic annihilation radiation.
These processes all occur after (but due to the nano- and picosecond timescales,
effectively parallel to) the primary and secondary processes and are denoted here as
post-effects. They are needed to finally understand the various characteristic emis-
sions, all induced by ✶K1, but caused by different pathways. This is discussed in the
following Chapter 12.
342 11 Secondary transformations
Tab. 11.7: Important γ-emitting(a) radionuclides originating from primary β−-, EC and
α-transformations and routinely applied for human diagnosis in nuclear medicine used for
scintigraphy or SPECT and for therapy, respectively.
https://doi.org/10.1515/9783110742725-012
344 12 Post-processes of primary and secondary transformations
pairs. However, there are some additional emissions inherently1 accompanying some
of the primary and secondary processes. These represent either particle radiation (var-
ious kinds of electrons) or electromagnetic radiation (X-rays and bremsstrahlung),
both of origin and characteristics not identical to the electron and photon emissions
released in primary and secondary processes. In principle, there are two classes of
post-effects occurring in condensed matter. Their origins are listed in Tab. 12.1.
Tab. 12.1: Types and origins of individual post-effects. Subtypes of primary and secondary
transformation processes yield either a positron, electron + positron pair or create a vacancy in the
electron shell of K2.
Primary β +
Ejection of a positron Annihilation radiation
from the atom (after thermalization)
The first type of post-effect relates to the appearance of a positron, which itself is
an antimatter particle and thus destined to annihilate its matter counterpart, the
electron (see Fig. 12.1). This necessarily happens for a primary β+-transformation of
a proton-rich nuclide, discussed in Chapter 8. Similarly, pair formation is a second-
ary transition in which a positron is created together with an electron. This occurs
via the conversion of electromagnetic energy iΔE > 1.022 MeV released between two
nuclear states (see Chapter 11).
In the second type of post-process, a vacancy (hole) in the electron shell of an
unstable nuclide is involved. The same hole can arise during a primary electron
capture transformation of a proton-rich nuclide, or during a secondary internal con-
version, as shown in Fig. 12.2.
These two principal effects stimulate two questions: What happens to the posi-
tron, and what happens to the electron vacancy?
These post-processes are inherently linked to the primary and secondary transformations of
the unstable nuclide ✶K1. The corresponding radiation emitted thus indicates the preceding nu-
clear pathways. However, post-effects that address a hole in an atom’s inner shell are also induced in
nonradioactive processes. Whenever an electron vacancy is achieved by external action (e.g. in X-ray
spectroscopy), the follow-up routes to fill these vacancies are the same.
12.1 Post-processes paralleling primary and secondary nuclear transformations 345
β+
e+
e
–
Fig. 12.1: Post-effects related to the appearance of a positron. This happens for a primary
β+-transformation and similarly for a secondary pair formation.
e
–
K
e
–
L
M
e
–
Fig. 12.2: Post-effects motivated by a vacancy in an electron shell of the atom. The vacancy arises
after a primary electron capture transformation of a proton-rich nuclide, or after a secondary
internal conversion.
346 12 Post-processes of primary and secondary transformations
The positron is the antimatter elementary particle of the electron with all the quan-
tum physical parameters of the electron.2 The exception is the charge, which is +1
instead of −1. Since antimatter vanished about 13.75⋅109 years ago,3 there are no
positrons “naturally” left in the universe. Yet, positrons appear in radioactive trans-
formations of many man-made unstable nuclides. The key question in the context
of post-processes is: What happens with the positron?
What happens with the positron? In a vacuum, the answer is: nothing! Its lifetime
is τ > 1020 a. This is different in (condensed) matter of course. Here (in contrast to
the electron neutrino), the positron intensely interacts via inelastic and elastic
scattering with the electrons located in the shell of atoms (atomic electron orbitals)
or molecules M (molecular electron orbitals). This interaction reduces the kinetic
energy of the moving positron until it is at almost zero: it has “thermalized”.4 The
total distance the positron travels depends on the kinetic energy it got from the
β+-event (remember this includes a continuous distribution between almost zero
and a maximum energy) and the kind of surrounding matter. It is particularly af-
fected by the Qβ + value, the values of Z and A, and the density of the chemical ele-
ments or compounds.
Like the electron neutrino, the positron was first predicted by P DIRAC in 1928 and only after-
wards experimentally verified (CD ANDERSON in 1933). DIRAC received the Nobel Prize in Physics in
1936 “for his discovery of the positron”.
The moment the universe was created within the Big Bang, matter and antimatter existed simul-
taneously in an (almost) 1:1 proportion. Within fractions of a second, however, the corresponding
matter and antimatter elementary particles disappeared by fusing with each other. This process is
called annihilation. There is no evidence of any antimatter in the universe, which would look like
an atom made of negatively charged protons in the nucleus and positively charged electrons (posi-
trons) in the shell.
“Thermalization” refers to the process in which a high energy object (such as a particle) interacts
with objects (atoms and molecules) of the surrounding environment. If those objects show (vibra-
tion) energy equivalent to room temperature, the higher-energy particles lose energy through inter-
action, and finally obtain the same energy as the surrounding particles. For particles at room
temperature, kinetic energy (of electrons) is Ee = 3/2 kBT (kB = BOLTZMANN constant, T = absolute
temperature) and follows a MAXWELL distribution. “Thermal energy” is equivalent to 0.025 eV.
12.2 Post-processes due to the positron 347
5 5
water
4 4
range (mm)
range (m)
3
air
2 2
Al
1 1
0 0
energy (MeV)
Fig. 12.3: Typical correlations between beta particle energy and “travel” distance in air, water and
aluminum.
Such correlations in terms of range of the beta particles in air have been used to correlate properties
of nuclides undergoing beta-transformations such as beta-particle energy or Qβ-values; see. eq. (8.17).
348
10 10
18F ( E max = 635 keV) 11 C ( E max = 970 keV)
5 5
0 0
-5 -5
4 mm
13 N ( E max = 1190 keV) 15 O (E max = 1720 keV)
5 5
0 0
-5 -5
4 mm
Fig. 12.4: Simulated tracks of positrons ejected from the decay of a point source of a sample of 100 nuclides of 18F (top) and spherical
distributions of ranges for prominent positron emitters (bottom). 18F (t½ = 110 min, Eβ+ = 635 keV), 11C (t½ = 20 min, Eβ+ = 970 keV),
13
N (t½ = 10 min, Eβ+ = 1190 keV) and 15O (t½ = 2 min, Eβ+ = 1720 keV). Average range distributions in water are mainly determined by the mean
kinetic energies of the positrons. They are in the order of 18F < 11C < 13N < 15O. See Fig. 8.7 for the distribution of kinetic energies of these
nuclides. With permission: CS Levin, EJ Hoffman, Calculation of positron range and its effect on the fundamental limit of positron emission
tomography system spatial resolution, Phys. Med. Biol. 44 (1999) 781–799.
12.2 Post-processes due to the positron 349
In the case of pair formation due to secondary processes, the positron and the
electron created annihilate immediately. In contrast, for positrons released in β+-
transformations, annihilation cannot occur directly due to the high kinetic energy
of the positron emitted. Here, this positron first interacts by scattering with sur-
rounding matter. Whenever inelastic scattering dominates, the positron stays
alive and travels further. Finally, the kinetic energy of the positron approaches its
zero value. Now, interaction of the positron with a shell electron becomes elastic
and finally the positron may combine with an electron. This represents the forma-
tion of a pair of matter + antimatter particles at about zero kinetic energy and re-
sults in the annihilation of the two particles.
The mass of the two particles converts into energy according to E = mec2. With the
(rest) mass of the electron of me = 9.109383⋅10−28 g = 0.00054858 u and the energy
equivalent of Ee = 510.9989 keV (see Table 2.1) the overall energy is 2 × 510.9989 keV =
1.0219978 MeV. This energy is emitted as electromagnetic radiation composed of two
photons of 510.9989 keV each, emitted in opposite directions, shown in Fig. 12.5. These
photons are measured γ-spectroscopically and appear in all the γ-spectra of unstable
nuclides undergoing primary β+-transformation or the secondary processes of pair for-
mation. Figure 12.6 shows a γ-spectrum of 18F.
Once inelastic scattering has reduced the kinetic energy of the positron to less than
1 keV, scattering becomes more and more elastic. With kinetic energies of about
<100 eV, the positron has – in addition to immediate annihilation – two more interac-
tion options. It can (1) form positronium Ps, and/or (2) ionize surrounding atoms and
molecules M by combining with the atom. These competing processes of molecular
ionization and/or positronium formation are schematically described by eqs. (12.2)
and (12.3). The other pathway of the positron is not to hit a shell electron, but to com-
bine with it. The Q-value for the formation of positronium depends on the ionization
potential UM of the counterpart electron of that molecule, eq. (12.4). This creates posi-
tively charged molecules, which remain for a very short period of time. Figure 12.7
summarizes the various process of a positron released through β+-transformation.
350 12 Post-processes of primary and secondary transformations
E = 511 keV
d ~ E β+
radionuclide
β+ e- 180°
E = 511 keV
Fig. 12.5: The positron close to rest energy combines with an electron to annihilate. The overall
mass of the positron and the electron is converted into two photons of 511 keV, which are emitted
in opposite directions (shown here two-dimensionally).
150000
18F
109.7 min
β+ 0.635
511 keV annihilation no
100000
counts per second
photons
18O
50000
10000
Compton scattering
0
0 200 200 600 800 1000 1200 1400 1600 1800 2000
-energy (keV)
Fig. 12.6: γ-spectrum of a positron emitter (18F). 511 keV annihilation photon emission is the
dominant peak due to ca. 96.9% β+-branching of the primary transformation 18F ! 18O (3.1% ε).
Other signals are low energy photons due to Compton scattering (see Chapter 1 of Volume II for
detection of radioactivity). Note the linear scale of the y-axis.
12.2 Post-processes due to the positron 351
direct
ANNIHILATION
(once thermalized)
IONIZATION
decreasing energy β+
β+ increasing e−
distance
e−
β+
subsequent to
formation ofthe Ps
Fig. 12.7: Schematic overview of the late stage processes of positrons close to zero kinetic energy
interacting with electrons, atoms or molecules. Direct annihilation is the most straightforward, but
the formation of positronium may occur as an intermediate. Close to thermalization, the positron
may also stick to the atom or molecule, forming positively charged intermediates e+M (M✶).
12.2.5 Positronium
The formation and fate of positronium needs a more detailed discussion. First, what is
positronium? The combination of the positron and electron may rest for a (very) short
while, before it continues to annihilate. This intermediate status is (like a chemical
element) called positronium, because it’s understood to be atom-like and similar to
hydrogen. It even has its own symbol: Ps. Figure 12.8 represents the structures of a
hydrogen atom, anti-hydrogen atom, and positronium. The negatively charged shell
electron is the same for H and Ps; the difference is the nucleus, with a positron for Ps
instead a proton for H. Of course, the masses of Ps and H are very different. And be-
cause of the much lower mass of the positron, the electron is less attracted in Ps.
This yields a larger radius for Ps and reduces its ionization potential.
Formation of positronium requires some kinetic energy of the positron to com-
pensate for the binding energy of Ps, which is EPs = 6.8 eV. In addition, one must con-
sider the ionization energy UM of the counterpart electron in the shell of the atom or
molecule. Thus, formation of Ps occurs at very low kinetic energy of the β+-particle (a
few eV, i.e. very close to rest energy).
352 12 Post-processes of primary and secondary transformations
e- e-
e- e+
p+ p- β+
β+
Fig. 12.8: The structure of positronium compared to the atoms of hydrogen and anti-hydrogen.
Bohr radius and ionization potential are identical for H and anti-H, but different for Ps.
Tab. 12.2: Overview of the properties of ortho – and para-Ps. For its quantum mechanics, physico-
chemical and nuclear parameters, see below.
0β 0e 0β 0e
+1 -1 +1 -1
Direction of transformation
condensed phase, however, produce a kind of quenching, and the mean lifetime is
much shorter, i.e. about 10−10 s. The basic equations are listed in Fig. 12.9.
Whereas the solid angle emission is stochastic for 3-photon annihilation, it is
almost exactly 180° for 2-photon annihilation.9 The effect of the simultaneous emis-
sion of two identical 511 keV photons in opposite directions is the basis of the so-
called coincidence measurement in positron annihilation processes and represents
one of the unique physical principles of Positron Emission Tomography (PET),10
shown in Fig. 12.20.
This angle may only differ from the ideally opposite direction if the positron deposited a small
amount of kinetic energy into the positronium being formed. In this case, relatively small devia-
tions of a few milliradians occur.
See Chapter 13 in Volume II.
354 12 Post-processes of primary and secondary transformations
πr2o 2 +4 +1 +3
(12.5)
= = [ ln( + 2 –1) ]
D 2
+1 √ 2 -1 2 -1
v«c
=1/ 1 –
(12.6)
D = 2 = πr2o –
v
3 1 3 3 (2J T +1) 1 1 1
(12.8)
(12.7)
= = = –=
2 1115 2 2 (2J s +1) 1155 3 372
Fig. 12.9: Formation probability σD and annihilation velocity λ of Ps, depending on singlet or triplet
quantum states. Γ is the velocity of the electron relative to the speed of the light; ro is the radius of
the electron. Transformation = annihilation constants are shown for 2-photon (λ2γ) and 3-photon
(λ3γ) emissions.
The first option to refill an electron vacancy proceeds through the import of elec-
trons populating higher shells of that atom. Figure 12.10 illustrates the pathways for
a vacancy in the K-shell or L-shell. Any electron hole is typically filled from an elec-
tron in the shell closest to it. Typically, an L-shell electron fills a K-shell electron
vacancy. In parallel, an electron from the N-shell may transit, but this process is
less abundant. Suppose the initial vacancy appeared in the L-shell; analogously,
electrons from the M – or N-shell may fill that hole. The transitions are named Kα or
Kβ for L ! K or M ! K, and Lα or Lβ for M ! L or N ! L routes, respectively. Here,
K or L indicate the position of the hole to be filled, and the Greek index indicates
12.3 Vacancies of shell electrons 355
e–
Kβ
e- e-
e–
K Lβ L
K K
e- e-
L L
M M
N
Fig. 12.10: Transition of electrons from higher shells to a vacancy in a K-shell (a) or L-shell (b) and
their notations.
whether the arriving electron descended from the next proximate (α) or the next but
one (β) main shell.11
This is analogous to the hydrogen emission spectrum. For the BALMER series, transition
for M ! L or N ! L would correspond to the notation of Lα = 656.3 nm or Lβ = 486.1 nm wavelength.
Because much heavier atoms are concerned for the post-effects, the ΔE values between the shells
are much lower and the electromagnetic emissions are not within the segment of visible light, as
per Fig. 11.5.
356 12 Post-processes of primary and secondary transformations
The relationship of type ν ≈ Z2 turns into ν½ ≈ Z with (Kα) = const (Z − 1) and ν½ (Kβ) =
const (Z − 7.4).
Furthermore, there are notations such as Kα1 and Kα2. This indicates electron
transitions into the K-shell vacancy from two different energetic levels within the L-
shell (due to the different quantum numbers l = 0 and 1). Their differences in ΔE are
very low, and their relative abundances are about 2:1.
!
hc 1 1
ΔE = h ν = = Ei − Ef = RH 2 − 2 Z2 (12:9)
λ ni nf
!
1 1
EKα = 13.6 2 − 2 Z2 ðeVÞ (12:10a)
1i 2f
!
1 1
ν Kα = 3.29 · 10 15
− Z2 ðHzÞ (12:10b)
12i 22f
!
1 1
Kβ = 13.6 2 − 2 Z2 ðeVÞ (12:11)
2i 3 f
Tab. 12.3: X-ray emission energies of selected chemical elements. Kα, Kβ, Lα, Lβ, etc. values
increase with increasing Z. Because ΔE is smaller for L ! K instead of M ! K transitions, Kβ-values
are larger by ca. 10% than the corresponding Kα-values of the same element [1]. Similarly, because
ΔE is smaller for M ! L compared to L ! K transitions, Lα-values are lower by a factor of ca. 10
than the corresponding Kα-values of the same element [2].
Z Kα Kβ Lα Lβ
Mg . .
Si . .
S . .
Ar . .
Ca . . . .
Ti . . . .
Cr . . . .
Fe
This contributed to the better arrangement of the chemical elements within the PTE not just
according to mass, which sometimes gives an irregular positioning, but clearly by Z. In initial struc-
tures of the PTE with elements arranged according to increasing mass, the sequence for the last
elements of the fifth period according to their mass would have been Sb (1 mol = 118.71 g) < I
(126.90 g) < Te (127.60 g) < Xe (131.29 g). This contradicted the idea of groups in the PTE containing
elements of similar chemistry, thus the correct order here should be Sb < Te < I < Xe. MOSELEY’s con-
tribution could ultimately clarify the situation, arranging the elements by proton number, i.e. 51Sb
< 52Te < 53I < 54Xe. Moreover, the correlations of type ν½ ≈ Z also confirmed the BOHR model of the
structure of electron shells. It was even used as an approach to search for chemical elements miss-
ing from the early PTE, such as 43 (technetium), 61 (promethium), 72 (hafnium), and 75 (rhenium).
In 1914 MOSELEY could predict their X-ray spectral lines. Accordingly, the two stable elements Hf
and Re were both discovered through X-ray spectroscopy. Both metals were chemically isolated
and their existence verified. In 1925, NODDACK, TACKE and BERG reported that they had detect ele-
ment 43 in platinum ore, in the mineral columbite, a niobium ore [(Fe,Mn)NbO6], in gadolinite and
in molybdenite through the X-ray signal of a wavelength characteristic for element 43. Because the
element itself could not be isolated chemically, the final proof remained open. Today the latter ap-
pears to be logical, as this (radio)element does not exist in measurable amounts – there is no stable
isotope of technetium!
358 12 Post-processes of primary and secondary transformations
Electron vacancies can also be refilled by a radiation-free process. The basic step re-
mains the same: transit of an electron from a nearby higher shell. However, ΔE is not
released as an X-ray, but is used to release another electron from a higher shell. The
process is referred to as a radiationless “reorganization” due to a “direct” interaction
of two electrons. This particular electron is immediately ejected from the atom; in
this case, no electromagnetic radiation is emitted. If electrons are emitted in the
course of processes between different main shells (interstate transitions), they are
called AUGER electrons (A). If the pathway involves subshells within the same main
shell, they are called intrastate transitions or COSTER–KRONIG electrons (CK).
10 6 207Bi
3.55 a
569.7 keV
K +K
207Pb
1063.7 keV
10 5
748.4 keV
897.8 keV
10 4
C ounts
1259.3 keV
10 3
K
K
10 2
10 1
EX-ray (keV) E (keV)
Fig. 12.11: Electromagnetic emissions of 207Bi: X-ray spectra and γ-photons. Primary EC and
β+-transformations of 207Bi (t½ = 31.55 a) induce secondary photon emissions (Chapter 11) and
post-effect characteristic Kα and Kβ lines. These are due to the transformation product nucleus:
stable 207Pb. The individual X-rays are indicated with their corresponding energy and intensity.
Courtesy: TS Vylov et al., Spektren der Strahlung radioaktiver Nuklide, gemessen mit
Halbleiterdetektoren, ZfK-399, Rossendorf, 1980.
example. The terminology is KLM. This fills the initial hole, but leaves two13 new
holes: one in the L-shell and one in the M-shell. The subsequent vacancies may
then be handled the same way, for example by using an M-shell electron to “refill”
the L-hole, thereby (for example) releasing an N-shell (A) electron (denoted as
LMN, Fig. 12.12(b)). (CK) electron pathways would be LLM, for example.
A complementary model of origin of (A) and (CK) electrons to explain their huge
number is drafted in Fig. 12.12(c). It gives a somewhat oversimplified representation
of cascade processes resulting in the release of a high number of (A) and (CK) elec-
trons. Suppose not just one, but two L-shell electrons start to simultaneously transit
to the initial K-shell vacancy; they would leave two new holes in their L-shell. Only
one is “accepted” in the initial K-shell vacancy, while the other is ejected from the
Remember that the X-ray cascade mechanism only induced one new electron vacancy for each
initial hole refilled.
360 12 Post-processes of primary and secondary transformations
a b
K K
L L
M M
N N
K
L
M
N
Fig. 12.12: Model of successive Auger electron emissions. (a) An initial vacancy in the K-shell is
filled by an L-shell electron, thereby causing a new vacancy in the L-shell. While the initial hole is
being filled, one M-shell electron leaves the atom (red line). The nomenclature of these emitted
Auger electrons is KLM. (b) The vacancy created in the L-shell is refilled analogously by e.g. an LMN
approach. Another Auger electron is emitted, this time from the N-shell. (c) Let us assume that not
only one L-shell electron identified the “energetically promising” vacancy in the K-shell, but two.
Only one would be successful in occupying the hole in the K-shell. The other, once departed from
its former L-shell, cannot return. At the same time, the two holes the two L-electrons left when
transiting to the K-shell vacancy are immediately identified by e.g. M-shell electrons. Let us
assume that now four of the M-shell electrons start to improve their binding energy, but again: just
two will be successful. All the electrons once released from their shells cannot return, because
electrons that transited from the next highest energy shell have now occupied their earlier
positions. They thus leave the atom. For the hypothetical model given, seven electrons have
already left the atom. They represent three individual shells energies (gray areas).
12.3 Vacancies of shell electrons 361
atom. At the same time, the two new holes created in the L-shell are identified by
four electrons of the M-shell. Similarly, all 4 start to transit, but again only two find
their destiny in the L-shell. The two others are released as (A) electrons. At this stage,
7 (A) electrons have already been ejected. Next, 8 N-shell electrons address the four
new holes in the M-shell (and so on). The process continues in a kind of cascade or
avalanche. Figure 12.13 shows the frequency of the overall (A) + (CK) electrons re-
leased per primary transformation step of 125I and their individual energies. Although
each electron has an individual discrete energy, the various electrons emitted within
several shell cascades represent a mixture of several characteristic energies, shown
in Fig. 12.13.
¹²⁵I
59.41 d
ε
; e–
¹²⁵Te
57.4 d
I (35)
e–
Rel. Frequency
ns
ctro
Ele
60
Nr.
40
20
0.01 0
0.1 1 10 100
Energy (keV)
Fig. 12.13: Auger and Coster–Kronig electron emission from 125I. Frequency of the overall (A) + (CK)
electrons released per primary transformation step of 125I and their individual energies. With
permission: DE Charlton, J Booz, KFA J 1979.
The overall number of electrons ejected is thus larger than the number of all the
shells (and subshells) of a given atom. Obviously, the number is larger in the case
of chemical elements of high main quantum number relative to those located in
lower quantum number periods. Table 12.4 summarizes the total number of AUGER
and COSTER–KRONIG electrons ejected for representative nuclides, as well as their
overall energy. It also compares this with the total number (and energy) of X-rays
simultaneously emitted within a single nuclide transformation. Indeed, the overall
number of (A) and (CK) electrons ejected increases with Z, but does not directly mir-
ror the number of atomic electron shells in the corresponding elements.
362 12 Post-processes of primary and secondary transformations
Tab. 12.4: Total number of Auger and Coster–Kronig electrons ejected and their overall energy
compared to the number and energy of X-rays simultaneously emitted for representative nuclides
(calculated). From: W Howell, Radiation spectra of Auger-electron emitting radionuclides: Report
No. 2 of AAPM Nuclear Medicine Task Group No.6. Medical Physics, 19 (1992) 1371.
After emission of all the shell electrons, what remains is a highly charged cat-
ion instead of a neutral atom. For 125I, for example, a nuclide, which primarily
transforms through electron capture to an excited state of 125Te, leaving a vacancy
in the K-shell, the number of AUGER and COSTER–KRONIG electrons emitted is about
25. This of course must cause dramatic changes in the chemical environment of the
newly created atom.14
Chemical consequences of processes like this are discussed in Chapter 7 in Volume II in the
context of radiochemical separations.
12.4 Example applications of post-effects 363
Typically, (A) and (CK) electron energies are approximately 20 to 500 eV. Their
range in aqueous phase (i.e. biological systems), for example, is short, about 1 to
10 nm.15
Once the vacancy of an s-shell electron has been handled, the radiative and radia-
tionless post-effects proceed in parallel, yet to different extents. The fluorescence
yield ωX-ray (ωK, ωL, . . . ) gives the percentage for the filling of an electron vacancy
through radiative processes. The total process of addressing the hole created by pri-
mary or secondary transformations of an unstable nuclide is then ωX-ray + ω(A)+(CK) =
1. The ratio between X-ray and AUGER and COSTER–KRONIG electron emission depends
on the proton number of the nucleus and thus the absolute energies of the electrons
and the relative differences between individual electron shell levels. Fluorescence
dominates at higher Z, which is for nuclides of the heavier elements. Figure 12.14 il-
lustrates this correlation in terms of the fluorescence yield for K-shell vacancies.
Yet the biological efficacy of (A) and (CK) electrons is high, because their kinetic energy is trans-
ferred to ionization processes over a very short distance; the Linear Energy Transfer is large (LET).
This becomes relevant in terms of endogenous radiation therapy, discussed in Volume II, Chapter 14.
The state-of-the-art in clinical applications of the corresponding radiopharmaceuticals for PET
is presented in Volume II: Modern Applications of Nuclear and Radiochemistry; specifically in
Chapter 12, entitled “Life Sciences: Nuclear Medicine Diagnosis”.
364 12 Post-processes of primary and secondary transformations
K 0.9
0.8
experimental
K-shell fluorescence yield
K
0.7 theoretical
K
0.6
0.5
0.4
0.3
0.2
0.1
0
0 20 40 60 80 100
Z
Fig. 12.14: Fluorescence yield for K-shell vacancies.
FDG is accumulated in cells of high glycolysis rate similar to glucose. Once it has entered the
cell, the first enzymatic degradation forms the 2-[18F]fluoro-2-deoxy-6-phosphate via the hexokinase
enzyme similar to endogenous glucose-6-phosphate. However, in contrast to glucose-6-phosphate,
the fluorinated 2-deoxy-6-phosphate of glucose is not further metabolized and remains trapped in
the cell. This results in an enhanced uptake of 18F in all cells of enhanced glucose turnover.
12.4 Example applications of post-effects 365
radionuclide
180°
0
+1 -1 e
Fig. 12.15: The principle of PET detection. Coincident registration of two 511 keV photons originating
from the annihilation of a positron + electron pair. Only those events are used for tomographic
measurements that are registered simultaneously (“co-incident”) in two opposite detectors (red)
within ca. 10 ns.
g glucose
ml . min
Fig. 12.16: [18F]FDG PET scan illustrating glucose metabolism in the brain of a healthy living human.
Normal glucose metabolism was seen in the 42-year-old volunteer (the author himself).
Single photon (or commonly γ-) emitters have been discussed in Chapter 11. Some
of these radionuclides are characterized by only one (or mainly one) γ-emission be-
tween different excited nuclear states. Generally, this transition is in high abun-
dance and it emits a relatively low energy photon (ca. 100–300 keV, ideal to be
registered by dedicated detectors). Such radionuclides should also have a clinically
and commercially convenient half-life, as well as an efficient production route. This
profile qualifies a few radioisotopes to function as radioisotope for SPECT, which is
single photon emission (sic!) computed tomography. Table 12.5 list the most promi-
nent candidates.18
Tab. 12.5: Important β+- and single photon (i.e. γ)-emitting radionuclides applied in nuclear
medicine used for PET and for SPECT, respectively. (✶) BNM-LNHB/CEA-Table de Radionucéides.
12.5 Outlook
In addition to the two post-effects discussed, there are some further effects that
cause radiative emissions. They are basically discussed in terms of the interaction
of radiation with condensed matter. One of these interactions should be mentioned
here, because it induces a kind of electromagnetic radiation that is observed in the
emission spectra recorded for a certain unstable nuclide. When an electron (liber-
ated through internal conversion, for example) passes the proximity of its atomic
nucleus, i.e. its strong electric field, it loses energy: The interaction leads to a loss
in kinetic energy (“to brake”). Due to the law of conservation of energy the energy
lost is converted into electromagnetic energy (in this case photons); referred to as
“bremsstrahlung”,20 as per Fig. 12.17. It is also called internal or “inner (or internal)
bremsstrahlung”.21
Its branching, however, is relatively low. The number of bremsstrahlung photons
per transformation is <10−3 (ΔE)2. Their energy is lower than the energy of X-rays
emitted for the same nuclide. A complete low-energy electromagnetic spectrum thus
may reveal both the characteristic X-rays and, at lower energy and with lower inten-
sity, continuous bremsstrahlung.
In addition, further effects arise from the interaction of the initial radiation caused
by primary, secondary and post-effect, i.e. both particulate and electromagnetic, with
condensed matter. However, these kinds of radiation interact with surrounding matter
in very different ways. Because this interaction is (in part) an inherent effect in
radiation measurement, these effects are discussed in Chapter 1 of Volume II in
more detail. It also becomes relevant in the context of how radiation interacts with
E´
ΔE = E´ - E´´ = hν
E´´
Fig. 12.17: The origin of bremsstrahlung. An electron originating from nuclear transformation
processes is attenuated by the nucleus of an atom. The kinetic energy the electron loses (from E´ to
E´´) is emitted as electromagnetic radiation.22
matter, especially its effects on biological material. The latter aspect relates to “radia-
tion dosimetry”, which is discussed separately in Chapter 2 of Volume II.
X-rays X-rays
ELECTROMAGNETIC EMISSIONS
particles particles
PARTICULAR EMISSIONS
IC electrons IC electrons
A + CK electrons A + CK electrons
radiation radiation
particles particles
neutrons neutrons
Fig. 12.18: Radioactive emissions structured according to their character, i.e. particulate vs.
electromagnetic (a) and according to the individual types (b).
This, actually, was the reason RUTHERFORD’s model of the atom was replaced by BOHR’s: the lat-
ter model allows electrons to “move” around the nucleus of an atom in quantum orbits exclusively;
see Section 1.3.3 (page 10).
370 12 Post-processes of primary and secondary transformations
bremsstrahlung particles
X-rays
particles
IC electrons
A + CK electrons
radiation
particles
neutrons
(a)
X-rays X-rays
particles particles
IC electrons IC electrons
A + CK electrons A + CK electrons
radiation radiation
particles particles
neutrons neutrons
(b) (c)
Fig. 12.19: Radioactive emissions structured according to their origin: emissions created in primary
transformations (a), secondary transitions (b) or post-effects (c).
This chapter has discussed the post-effects either related to “defects” created by pri-
mary and secondary transformations within the newly created nucleus K2 and its
atom, or to the formation of an antimatter particle within that nuclide, the positron.
The corresponding radiation generated by post-effects is either electromagnetic (X-rays,
annihilation photons and/or low energy inner bremsstrahlung) or particulate (cas-
cades of low-energetic AUGER and COSTER–KRONIG electrons). These radiations complete
the spectrum of primary (β−- and β+-particles, α-particles, neutrons) and secondary
(photons, IC electrons, electron-positron pair) emissions.
Only together do all of these effects reflect the essence of radioactive transfor-
mation processes.
12.5 Outlook 371
Figures 12.18 and 12.19 group the various radiations by type, i.e. particulate vs.
electromagnetic (Fig. 12.18(a)), according to the specific transformations (Fig. 12.18(b)),
and according to the origin of the emissions created in primary transformations, sec-
ondary transformations and post-effects (Fig. 12.19). These radiative emissions are
all part of nuclear and radiochemistry – they all are constituting what is referred to
as RADIOACTIVITY.
13 Nuclear reactions
Aim: Unstable nuclei convert into stable ones by minimizing the absolute mass of the
nuclide, which is typically expressed in terms of nucleon binding energy. The com-
mon feature of all these transformations is their exothermic character, with transfor-
mations occuring spontaneously. In contrast, nuclear reactions start via external
“activation”, delivered by a projectile. Balances in the mass of reactants and products
may differ between exothermic and endothermic routes, and in the latter case the
projectile delivers activation in the form of kinetic energy.
The general strategy of artificial nuclear transformation is to induce a nuclear re-
action by delivering energy (either in the form of a particle or electromagnetic) of ade-
quate kinetic energy into a (mainly) stable target atom. Thereby, a compound nucleus
is formed, which instantaneously transforms into the nuclear reaction product(s).
There are several pathways for nuclear reactions and several classes may be de-
fined. Some occur naturally in the universe, particularly in stars or interstellar
space, and some occur in the earth’s atmosphere. Another group is comprised of
artificial, man-made nuclear reactions.
This chapter focuses on the latter class, and on the radionuclide product of the
reaction. It discusses the principal parameters influencing the differential and inte-
gral yields of radionuclides produced in certain nuclear reactions. These parameters
are: the cross-section and excitation function, particle energy and flux, stopping
power, irradiation period, and target composition.
Radionuclides have been and are of great value to basic research, for analytical studies
in astrophysics and environmental research, in chemistry and materials sciences, in
industry, and for medical diagnoses and therapy. However, access to radionuclides oc-
curring naturally on earth is limited to the members of the transformation families
starting from 232Th, 238U, and 235U contained in ores in the earth’s crust, and, at least in
principle, to a very limited number and in a very limited way to radionuclides perma-
nently created in the atmosphere. Consequently, it is the task of man-made nuclear re-
actions to produce “artificial” unstable nuclides, i.e. those not usually found on earth.
A nuclear reaction is, in analogy the chemical reaction, defined as the interaction
of (in almost all cases) two reactants to yield the desired reaction product(s). If a ra-
dioactive nuclide is to be produced, one reactant is typically a nucleus of a stable
atom (the “target”). The other is either a subatomic particle (mainly neutrons and
small charged particles like protons, deuterons, α-particles, etc.) or the nucleus of a
medium-large atom (electromagnetic radiation can also be used.). Thus, the initial nu-
cleus is “bombarded” or “irradiated” by the small “projectile” of characteristic kinetic
https://doi.org/10.1515/9783110742725-013
374 13 Nuclear reactions
energy and is transformed into a new nucleus. Efficient production, in terms of high
yields and high purity of the desired radionuclide, requires the following:
– adequate choice of the type of the nuclear reaction,
– availability of the corresponding projectiles and targets,
– handling the irradiation process, and
– (in many cases) the (radio)chemical separation of the radionuclide produced
from the irradiated target.
This is the search for the “philosopher’s stone” (lat.: lapis philosophorum; arab.: el Iksir). Its dis-
covery was said to be the magnum opus. Despite laborious efforts over centuries, it was never
found, although chemistry is thankful for the discovery of many initially unrevealed properties of
chemical substances. A fortunate spin-off effect was the contribution of JF BÖTTGER in 1707 on how
to prepare European porcelain. At the time, he had been hired by the monarch of Saxony in Dres-
den to convert less noble metals into gold.
13.1 Artificially produced radionuclides 375
In both cases, the nuclear reaction product formed was stable and thus “invisible”
in those reactions – but even then, these products were already known, stable iso-
topes (see Tab. 13.1).
In contrast, when I CURIE and F JOLIOT in 1934 used the same α-particles to irra-
diate natural elements such as boron, magnesium or aluminum,2 the irradiated
samples became radioactive (see Tab. 13.1). As chemists, they were able to chemi-
cally separate the nuclear reaction product (e.g. 13N as ammonia from the irradiated
boron nitride) and thereby could identify the elemental identity of the radioisotope.
This was the first artificial creation of unstable, “man-made” radionuclides, deemed
“artificial” because these nuclides do not exist naturally on earth!3
Tab. 13.1: First man-made nuclear reactions. Based on α-projectiles obtained from, for example, the
primary transformation of natural 226Ra yielded either stable or unstable (i.e. radioactive) product
nuclei. Half-lives for the latter were t½ = 9.96 min for 13N and t½ = 2.50 min for 30P.
The natural isotopic compositions of the irradiated stable isotopes (in today’s nomenclature,
“targets”, see below) are: 10B (19.9%) + 11B (80.1%) for boron, 24Mg (78.99%) + 25Mg (10.00%)
+ 26Mg (11.01%) for magnesium, 27Al (100%). FRÉDÉRIC JOLIOT reported: “We have proposed that
these new radio-elements (isotopes, not found in nature, of known elements) be called radio-
nitrogen, radio-phosphorus, radio-aluminum (in the case of magnesium irradiated by alpha rays)
and designated by the symbols: R13N, R30P, R28Al”. Note that an (α,n) reaction on one of the magne-
sium isotopes would not yield an isotope of aluminum, but of silicon. The isotope 28Al (t½ =
2.246 min) could be formed via a 25Mg(α,p) process.
The Nobel Prize in Chemistry, 1935, was awarded jointly to FRÉDÉRIC JOLIOT and IRÈNE JOLIOT-CURIE
in recognition “ . . . of their synthesis of new radioactive elements . . . ”. The Nobel lecture was
entitled: “Chemical Evidence of the Transmutation of Elements”; note Transmutation of Elements.
This was organized through α-particles of relatively low energy, which originated from naturally
occurring radioactive sources. Another Nobel Prize was awarded for “the transmutation of atomic -
nuclei” using artificially created charged particles, as per footnote 37.
376 13 Nuclear reactions
13.2.1 Nomenclature
The course of a nuclear reaction is typically written as if a small component “a” deliv-
ers mass and/or energy to a larger nucleus “A”, forming a product nucleus “B” and
eventually a small component “b”, as in eq. (13.1): a + A = B + b. The short notation is
A(a,b)B given in eq. (13.2). The nucleus A is called the “target”. The small component
a is the “projectile”, and b is the “ejectile”. The nuclear reaction procedure is often
said to “irradiate” a target, and the target is “irradiated” or “bombarded”.
a+A=B+b (13:1)
One may further distinguish between naturally occurring and man-made nuclear
reactions. Some nuclear reactions take place naturally in the universe, in stars
(such as the fusion of light elements and genesis of heavier elements) or in super-
novae (such as the permanent nuclear synthesis of 26Al in supernovae, as men-
tioned in Chapter 11). Some naturally occurring nuclear reactions take place in the
earth’s atmosphere (such as formation of 14C).
Artificial, man-made nuclear reactions govern the direct interaction of projectiles,
either naturally available or artificially produced, with dedicated target nuclei.
In the context of “particle research”, nuclear reactions aim at infinitesimal ele-
mentary particles and their properties4 and at the origin of the infinite universe and
astrophysics in general. Another group may be identified revealing the mechanism
of nuclear reactions itself, which is basic physics again.
In the framework of nuclear and radiochemistry, however, there is a specific
goal: The production of artificial radionuclides for
– chemical and physical research and development,
– research on nuclear materials,
– industrial analytics, and
– the synthesis and application of radiopharmaceuticals in medicine.
A + a ! C✶ ! B + b (13:5)
For example, in 2012, CERN, Geneva (Conseil Européen pour la Recherche Nucléaire) confirmed
the existence of the HIGGS boson – see Fig. 7.1.
The total excitation energy may be expressed in temperature scale: C✶ is a “hot” nucleus. Parts of
this excitation are released in a prompt way, typically in terms of electromagnetic radiation. In par-
allel, nucleons or nucleon clusters may “evaporate” or “distill off” – like molecules above their
boiling point in “hot” solvents.
378 13 Nuclear reactions
In nuclear reactions, fundamental parameters such as the sum of reactants and the
sum of products must remain the same. The intermediate compound nucleus C✶
thus obtains a nucleon composition that is exactly the sum of the nucleon num-
bers of a + A. For the total mass number A, this gives AA + Aa = AC✶ and AC✶ = AB
+ Ab. This is the same for each type of nucleon; for example, the number of neu-
trons is nA + na = nC✶ and nC✶ = nB + nb. This holds true for mass m of the partici-
pating species, and thus also for energy as E = mc2. The same conservation holds
true for other parameters such as the electric charge, momentum (linear and an-
gular) and parity.
AA + Aa = AC✶ = AB + Ab (13:6a)
mA + ma = mc✶ = mB + mb (13:6b)
Fig. 13.1: Mechanism of nuclear reactions, including an intermediate compound nucleus C✶.
This composite state C✶ splits instantaneously (typically with a lifetime of about 10−13 s)
into B and b. The mass of component b ejected from C✶ is typically much less than that
of the radionuclide B formed. It thus may leave the process at high velocity. In many
cases, the ejectile b is a photon γ, but in other cases can be a neutron, proton, deu-
teron, triton or α-particle. Whenever, for example, a neutron is released from the com-
pound nucleus, it removes about 8 MeV from the excitation energy of C✶ – which is the
mean binding energy of a nucleon. The second species formed is the nucleus B. If un-
stable, i.e. radioactive, it immediately starts to transform according to the processes de-
scribed in Chapters 8–12. It is this radiation that makes the product radionuclide B
valuable for specific applications!
13.2 General mechanism of nuclear reactions 379
Any given composite C✶ (as defined by Z, N, and A) may be obtained through different
combinations of type A + a, all having the same sum of protons and sum of neutrons.
The processes of forming a given C✶ are called “in-channels”, shown in Fig. 13.2.
A1 + a1 B1 + b1
A2 + a2 C* B 2 + b2
A3 + a3 B3 + b3
Fig. 13.2: Mechanism of nuclear reactions including nucleon transfer processes. A compound
nucleus C✶ defined by a given (Z, N, and A) may be obtained through various combinations of type
A + a, all representing the same sum of protons and the same sum of neutrons. These processes of
forming C✶ are called “in-channels”. From C✶, various “exit-channels” may open, depending on the
energy deposited into C✶.
The same is true for the fate of compound nucleus C✶. This energetically rich inter-
mediate state is transformed into the products B and b. Again, different combinations
of B + b may exist, reflecting the same sum of proton number, neutron number, and
mass number as C✶. These are the “exit-channels” (or “out-channels”).
A variety of in- and exit-channels are conceivable for a given compound nu-
cleus – all candidate channels must, however, be permitted by energy and conser-
vation laws. Why one particular in- or exit-channel is favored over others is decided
quantum mechanically, in terms of the degree of overlap between nuclear wave
functions of initial and final states.
Figure 13.4 summarizes the most commonly used nuclear reactions in producing ar-
tificial radionuclides with the projectiles γ, n, p, d, 3He, and α. The Chart of Nu-
clides indicates the stable target nuclide and the various product nuclei according
to the (a,b) nuclear reaction.
380 13 Nuclear reactions
15N + 4
7 8 2 2
17O + 21d1
8 9
13N + 1n
7 6 0 1
14N*
10B + 4 7 7
5 5 2 2
13C + 1p
6 7 1 0
Fig. 13.3: Examples of in-channels and exit-channels. The most relevant channels are highlighted.
The example for in-channel options is on the composite nucleus 19F✶, which leads to the positron
emitter 18F through an 18O(p,n) 18F reaction. For the exit-channels, an example for one of the first
man-made nuclear reactions is selected, namely 10B(α,n) 13N via the 14N✶ composite nucleus
formation. The in-channel was the one initiated by α-particles of relatively low kinetic energy
(offered by α-emitting members of naturally occurring chains), the exit-channel creating 13N was
the one energetically favored under these conditions.6
The energy balance in nuclear reactions based on nucleon transfer parallels a funda-
mental rule known in chemical reactions: stoichiometry. Yet, there is a very important
For the discussion of which in-channel works best and which exit-channel is preferred, see
below (energetics of nuclear reactions). The other exit-channels may appear with the increasing ki-
netic energy of the α-projectile. Possible exit-channels are [14N✶] = 13N + 1n and [14N✶] = 13C
+ 1p. The latter exit-channel “opens” through a higher kinetic energy E of the incoming α-particle.
The higher energy deposited in the composite nucleus may also allow for the separation of not just
one, but two nucleons.
13.2 General mechanism of nuclear reactions 381
(d,t)
(n,2n) TARGET (n, )
NUCLEUS (t,p)
( ,n)
n) (d,p)
(d p)
(p,d)
(n,d) (n,p)
(p, ) (d, )
( ,p) p
(d,2p)
PROJECTILES:
,
n, (n, )
p, d, 3He,
Fig. 13.4: Typical nuclear reactions used to produce radionuclides with the projectiles γ, n, p, d
3
He, and α within the Chart of Nuclides. The stable target nuclide is black, and the various product
nuclei are denoted with the balance of (a,b)-type nuclear reaction. For example, the neutron
capture reaction (n,γ) yields a nuclide just to the right of the stable target nuclide.
The latter reaction has an equilibrium constant KEQU = 4 at 800 °C, which corresponds to 80%
reaction yield of CO2 (and still containing 20% CO). By starting from 100% CO, this concentration
decreases while CO2 is simultaneously formed. This equilibrium is reached at a certain time, as
shown in Fig. 13.5, left.
382 13 Nuclear reactions
k + ½Aα A ½aα a
K= = (13:7)
k − ½Bα B ½bα b
13.3 Energetics
Then each α-particle had reacted a 10B atom into 13N, cumulatively yielding 1010 atoms of radio-
active 13N. With a half-life of t½ = 9.96 min (and supposing that the irradiation was arranged
within a very short period of time relative to this half-life in order to avoid the need to correct the
radioactivity lost by simultaneous transformation), the radioactivity of 1010 atoms of 13N would be
11.6 MBq. (Á = λ Ń = (ln2 / t½) Ń = (0.693 / 9.96 × 60 s)⋅1010 = 0.00116 s−1⋅1010 = 11.6⋅106 s−1 =
11.6 MBq.) This activity is significant, yet the changes in atoms of the target atom 10B are negligible.
F JOLIOT in his Nobel lecture, entitled “Chemical Evidence of the Transmutation of Elements”:
“The yield of these transmutations is very small, and the weights of elements formed by using the
most intense sources of projectiles which we are able to produce at the present time are less than
10−15 g, representing at the most a few million atoms.”
13.3 Energetics 383
100 A
80
%)
B B
ntrations (%
60 radioactivity
of the product nucleus
(arbitrary scale)
40
concen
20
0 B
0 1 2 3 4 5 0 1 2 3 4 5
Fig. 13.5: Comparison of chemical and nuclear reactions. Conventional chemical equilibrium of
CO + H2O ! CO2 + H2 with CO = A and CO2 = B, defined by the simultaneous decrease in A and
increase in B (left). In contrast, nuclear reactions do not result in a significant decrease in the
concentration of A, because the number of product nuclei B is comparably small (right). However,
this number reflects a significant radioactivity of B (dotted line).
13.3.2 Q-values
endothermic or exothermic
Fig. 13.6: Interplay of nuclear processes. Creation of an artificial radioactive nuclear reaction
product B (here: nucleus ✶K2) by inducing a process on a stable nuclide A (here: nucleus K1) and
subsequent exothermic nuclear transformation of the unstable product nuclide into a stable
nucleon configuration K3 (K3 may in some cases be identical to K1).10
A + a ! B + b + ΔE (13:9)
For example, the nuclear reaction of type 14N(n,p)14C, i.e. the naturally occurring pro-
cess to permanently produce 14C in the atmosphere of the earth, is 14N + n ! 14C + p
+ Q. Mass-wise, 14C plus one proton is lighter than 14N plus one neutron. The process
is exothermic, with a Q-value of −(−0.000672 u)11 = 626 keV.12 Actually, all (n,γ)
and (p,γ) type nuclear reactions are exothermic. Nuclear reactions induced by a
charged projectile larger than the proton can be either exothermic or endothermic.
While eqs. (13.10) and (13.11) use absolute mass, the mass excess values Δexcess
(or simply Δ, see Chapter 2) can also be applied as tabulated in the atomic mass
evaluation database (see also Table 14.4). The analogous equation to (13.11a) is:
Q = − ΔE = ΔA + Δa − ΔB − Δb (13:11b)
See for example the process of producing 18F from Fig. 13.3. The nuclear reaction is 18O(p,n)18F,
and the primary transformation of the proton-rich 18F is β+-emission to stable 18O.
Δm = [m(14N) + m(1n)] – [m(14C) + m(1p)] = (14.003074 u + 1.008665 u) – (14.003242 u + 1.007825 u) =
(15.011 739–15.011 067) u = − 0.000672 u. (1 u c² = 931.49 MeV).
Similar to nuclear transformation processes, this energy is mainly “located within” the new nu-
cleus in terms of nucleon binding energy.
13.3 Energetics 385
The first man-made synthesis of an artificial unstable nuclide was 27Al(α,n)30P. The
atomic masses are 26.981538 u for 27Al and 29.978313 u for 30P. With atomic mass of
the α-projectile of 4.002603 u and 1.008665 u for the neutron ejected, the balance for
the A(a,b)B nuclear reaction according to eq. (13.10) is Δm = 30.986978 u - 30.984141
u = 0.002837 u. Though this is a small change in atomic mass units relative to the
masses of (A + a), just 0.00915%, it is endothermic: the sum of the masses of the two
products is larger than the sum of the two reactants. When expressed as energy, it
gives a Q-value of −2.644 MeV.13 The reaction may thus only work if the α-particle de-
livers a kinetic energy above that level.14 For projectiles with a certain kinetic energy,
this kinetic energy (Ekin = ½mv2) must be added to eq. (13.11). If this referred exclu-
sively to the incoming projectile, this would result in eq. (13.12). In fact, this is the
main feature of most of the man-made nuclear reactions to produce a desired artificial
nuclear reaction product: to induce the reaction by delivering energy15 (either in the
form of a particle or electromagnetic radiation) to a target nucleus. In most cases of
radionuclide production, the target is at rest energy.16
1
EðA + aÞ = mA c2 + ma c2 + ma v2 (13:12)
2
In addition, the product species also receive an impulse, and thus the complete bal-
ance in energy is even more complex. Both nuclear reaction products are thus distrib-
uted along a virtual axis between projectile and target in an angular distribution. The
experimental measurement of these angles represents a valuable approach to deter-
mine the characteristics of a nuclear reaction process.
27Al: EB = 224.952 MeV, ĒB = 8.332 MeV; 30P: EB = 250.605 MeV, ĒB = 8.354 MeV.
Because this was in fact the case with the kinetic energies of α-particles emitted from the long-
lived α-sources of 226Ra and its daughter nuclides (4.784 MeV and 4.601 MeV for 226Ra, 5.489 MeV
for 222Rn, 6.002 MeV for 218Po etc.), this endothermic process took place.
Here is another difference between chemical and nuclear reactions – the absolute scale of the
energy involved. Exchanges between electrons of different species to form a molecule shift atomic
electrons to molecular electron levels. This proceeds on the scale of electron binding energies, i.e.
in the order of eV to keV. In nuclear reactions, nucleon binding energies are concerned, and “acti-
vation” energies may reach MeV scales. In modern particle physics, nuclear reactions are induced
on the level of GeV or even TeV.
For most of the commonly organized nuclear reactions, the target nucleus represents a stable
nuclide and is mechanically fixed. All the stable isotopes of the chemical elements can be used as
targets, either in the chemical form of the atom or in chemical compounds including that atom. In
some cases, such as the synthesis of super-heavy elements (see Volume II, Chapter 10), radioactive
target nuclei are also used.
386 13 Nuclear reactions
Exothermic nuclear reactions may occur at any kinetic energy of the projectile,
even (hypothetically) at its rest mass energy (i.e. projectile kinetic energy = 0). This
is different in endothermic nuclear reactions. The corresponding Q-value represents
a kind of energy that must be overcome by adding kinetic energy to the rest mass of
the projectile – it serves as a kind of activation or excitation energy E✶ for the reac-
tion itself: E✶ = −Q. The amount of E✶ is the excess energy deposited within the com-
pound nucleus C✶.
However, the kinetic energy of a projectile is not 100% used. Some part of this
energy is not available for inducing the nuclear reaction, because the projectile and
the target nucleus also obey conservation of impulse. The product nucleus gets a
recoil energy RECOILEB, which depends on the masses of both a and A, as per eq.
(13.13) (see also Fig. 9.2). Consequently, the kinetic energy of the projectile must be
“a bit” larger than the value of −Q in order to compensate for the recoil energy. To
be precise, it must be larger by −Q(ma/mA). The kinetic energy derived represents
the “threshold energy” Ěa of a nuclear reaction, shown in eq. (13.14):
ma
RECOIL
EB = E✶ (13:13)
ma + mA
mA + ma ma
Ěa = − Q = −Q 1+ (13:14)
mA mA
RUTHERFORD‘s nuclear reaction 14N(α,p)17O, for example, has a Q-value of −1.19 MeV.
The threshold energy is deduced via the ratio of the mass numbers of the α-particle
and the 14N nucleus: Ěα = −(−1.19 MeV) (4 + 14) / 14 = 1.19 MeV⋅1.2857 = 1.53 MeV.
The threshold energy in this case is larger than the Q-value by 0.34 MeV, i.e. about
28.5%. Nevertheless, the reactions could proceed because the α-particles kinetic en-
ergies were >4 MeV. For nuclear reactions of type (p, n) on target nuclei of mass
number 100, for example, the additional amount of kinetic energy of the protons
needed is only (1 + 100) / 100 = 1% of the Q-value.
Both the activation energy E✶ and the threshold energy Ěa discussed above exclu-
sively referred to the mass of the participating nuclear reaction components.
The nucleons of the projectile will benefit from the strong force if they are close
enough to the central attractive well of the target nucleus, which is about the di-
mension of a nucleon. However, target nuclei do have a positive charge. Thus, pro-
jectiles of positive charge like protons and α-particles (actually, all projectiles
13.3 Energetics 387
except for photons and neutrons) must overcome a repulsive coulomb force – much
longer ranging than the attractive strong force. Figure 13.7 compares the difference
between a neutral projectile (a neutron), and a positively charged projectile (a pro-
ton), in terms of approaching the target nucleus.
n p
Potential
Nucleus
well
with radius
r
Fig. 13.7: Projectiles entering the potential well of a target nucleus. Coulomb barriers apply to
charged projectiles, such as the proton, but not to uncharged projectiles, such as the neutron.
Note: Coulomb barriers must be taken into account both when entering the target nucleus and
when leaving it, such as discussed for primary transformations of unstable nuclei through
α-particle emission.
The coulomb barrier applies only to charged projectiles, not to noncharged projec-
tiles. The former need to have sufficient kinetic energy in addition to the threshold
energy, as in Fig. 13.8. The potential is U = f(r), with rC being the separation distance
between the (centers of charge of the) two components A and a. A maximum value is
obtained for the closest distance rC between them, which is given by eq. (13.15); ro is
the radius parameter, discussed in Chapter 2. It yields the value of the coulomb bar-
rier EC for the corresponding components, as in eq. (13.15) or (13.16). Suppose the ki-
netic energy of a charged projectile is below this level; it will not be able to reach the
attractive strong forces – it just may scatter elastically from the target nucleus.
Z A · Za · e 2
EC = (13:15)
rC
Z A · Za
EC ≈ 1.44 MeV fm (13:16)
rC
388 13 Nuclear reactions
1=3
rC = r0 AA + A1=3
a (13:17)
1/r
U(r)
EC
U(r)
potential well
rC
rNUCLEUS
radius r
Following the intention to artificially produce radioactive nuclides, either for basic
research or for industrial and medical applications, nuclear reactions should be ex-
perimentally designed in a way to achieve maximum product nuclide yields (activi-
ties) and highest product purities (similar to conventional preparative chemistry).
Again, similar to conventional chemistry, fundamental correlations exist, defining
nuclear reaction yields. These parameters are discussed below.
Suppose a projectile has sufficient kinetic energy to enter the target nucleus; by far,
not every hit represents a nuclear reaction. Usually, many “hits” are needed to in-
duce just one nuclear reaction. The ratio of the number of hits to the number of
13.4 Yields of nuclear reaction products 389
A1 A1 = A2
= r2
with rnucleus = 1.10-14 m
Anucleus = 1.10-28 m2 = 1 barn
Fig. 13.9: Intention of the “nuclear reaction cross-section”. Schematic representation of the
different probabilities for a projectile to hit a small and a large target. For the nucleus on the right,
the probability of a bullet hitting the target is twice as high as for nucleus on the left: two out of
two projectiles reached the target as opposed to just one. For the same bullet, the nuclear reaction
“cross-section” is small for the target nucleus on the left and large for the one on the right. Note
that the gray area on the right is virtual – the real nucleus (black) does not change its dimension.
How does this apply to nuclear reactions? The target nucleus is taken with its real
dimension, which is on average rnucleus ≈ 10−14 m. Its diameter represents a plate or
a circle, whose area is about the square of rnucleus, i.e. of Anucleus ≈ 10−28 m2. This
unit is called 1 barn (b) = 1 = 1⋅10−28 m2, which is extremely small.17 This area
stands perpendicular to the arriving projectile.
This is a joke; “barn” in English refers to a barn door, which is big enough to let horses and
farm tractors pass.
390 13 Nuclear reactions
From a simple geometric point of view, the same number of identical projectiles
should then yield the same nuclear reaction rate for target nuclides of the same (or
very similar) mass number. However, this is not even remotely true. Figure 13.10 gives
three stable target nuclides of same (isobar A = 168) or similar mass (A = 169). However,
the probabilities for a neutron to induce a nuclear reaction (neutron capture, (n, γ))
differ by three orders of magnitude. For the incoming neutron, the 169Er nucleus
“looks” like a “normal” nucleus with ca. 2⋅10−28 m2 area (σ = 2.3 b), but the 169Tm nu-
cleus appears to be about 50 times larger (σ = 108 b), and the 168Yb nucleus about 1000
time larger (σ = 2400 b)! All three nuclei are almost identical in size, but the probability
of incorporating a neutron is about 1000 times higher for 168Yb than for 169Er.
168Yb 169Yb
0.13 32.0 d
2400
169Tm 170Tm
128.6 d
100
108
168Er 169Er
26.978 9.40 d
2.3
Fig. 13.10: Neutron capture cross-sections σ for AA(n, γ)A+1B reactions on lanthanide nuclides of
identical or similar mass number A. Values are quite different, i.e. 2.3, 108 and 2400 barn for the
168
Er(n, γ)169Er, 169Tm(n, γ)170Tm and 168Er(n, γ)169Er, respectively. (Italic numbers give the
abundance H in % of the stable isotope for the natural isotope composition of the element.).
This is because the very simplified geometric explanation has a quantum mechani-
cal background. The nuclear reaction probability Wif is bridged with the overlap of
initial and final wave functions,18 and the cross-section finally mirrors the transi-
tion matrix element (see Tab. 13.2). The probability of a certain nuclear reaction is
high or low because of quantum mechanical considerations. The cross-section
serves as a proportionality constant. If – for the same number of projectiles, given
in units of 1/m2 – the probability of a certain nuclear reaction is high, then the
. . . analog to the probability Pfi of transition (per unit time) between initial (i) and final (f)
states of unstable nuclides undergoing radioactive transformation, in particular for primary β-
transformation, as shown in eq. (7.2). For radioactive transformation, quantum mechanics defined
a transition rate that parallels the transformation constant or half-life of the processes.
13.4 Yields of nuclear reaction products 391
cross-section value is high and vice versa. In this respect, the cross-section is similar
to the equilibrium constant used in conventional chemistry. There are even ex-
tremely low probabilities, expressed as σ < 1 (see Fig. 13.11) indicating a cross sec-
tion much smaller than the area of the target atom!19
Tab. 13.2: Quantum mechanical background of the cross-section. Mif = transition matrix element,
Eif = density of energy states, ρpsf = phase space factor.
Probability Wif = reaction rate (events/s) Wif = π/ħ r(Eif) {Mif} = ρpsf {Mif}
39Ar
269 a
3
2
37Cl
0.8
Fig. 13.11: Cross-section values σ (in barn) for different neutron-induced reactions on stable
potassium target nuclei. The 39K(n, γ)40K process has a cross-section of σ = 2.1 b, that of the
41
K(n, γ)42K process is σ = 1.46 b [1]. For 39K, the 39K(n,α)42Sc (σ = 0.0043 b) [2] and the 39K(n,p)39Ca
(σ < 0.0005 b) [3] processes are much less probable.
Cross-section values cover a very broad range, from a few mb (1 mb = 10−27 cm2) to
105 b. Values (in barn) for neutron capture processes are indicated in the Chart of
Nuclides, shown in Figs. 13.10 and 13.11. If not indicated in detail, the values stand
for (n,γ) processes induced by neutrons of thermal energy, which is 0.025 eV (up
to 0.1 eV). For some special cases, cross-section values of (n,γ) processes induced
The σ-value of the 4He (n,γ) reaction is practically zero, although the area of the 4He nucleus is
not, of course.
392 13 Nuclear reactions
by fast neutrons (10 keV to 20 MeV kinetic energy) and for (n,α) and (n,p) reactions
are listed, such as 39K in Fig. 13.11.
scattering
exit-channel 1
total exit-channel 2
nucleon transfer
exit-channel 3
absorption spallation
fission
Fig. 13.12: Individual vs. total cross-section. While the total cross-section covers all the possible
interactions of the projectile with the target nucleus, the specific route yielding the desired nuclear
reaction product (here exit-channel no. 2 of a nucleon transfer reaction) just represents a part of
the total cross-section, an individual one.
At a given excitation energy, an exit-channel opens for the compound nucleus. The
“easiest” way to de-excite is to release electromagnetic radiation: in doing so, the
(n,γ) or (p,γ) nuclear reaction occurs. At very low excitation energy of C✶, the re-
lease of one single nucleon or cluster (e.g. the α-particle) is possible. The release of
several individual nucleons may only occur at higher excitation energies.
If the excitation energy is high enough, several exit-channels are possible from
the point of view of energy. For a proton-induced process, for example, these are
(p,γ), (p,n), (p,2n), etc., (p,pn) etc., (p,d), (p,α), etc. Figure 13.13 illustrates some of
13.4 Yields of nuclear reaction products 393
COMPOUND
NUCLEUS
+p
(p,pn)
TARGET
NUCLEUS
(p, )
PROJECTILE = PROTON
Fig. 13.13: Some exit-channel pathways for a proton-induced nuclear reaction within the Chart of
Nuclides. Target nucleus A, compound nucleus C✶, and projectile a according to the A(a,b)B,
i.e. A(p,b)B processes.
these exit-channel pathways and the position of the product nuclei relative to the
target nuclide and the composite nuclide.
The question is, which exit-channel works best? In fact, a certain individual exit-
channel may dominate over others. For a given combination of target nucleus and pro-
jectile, each exit-channel is individually populated depending on the kinetic energy of
the projectile; that is, on the excitation energy of the composite nucleus. This correla-
tion of type σ = f(Eprojectile) is called the “excitation function”. A typical excitation func-
tion for a nuclear reaction induced by charged projectiles is given in Fig. 13.14,
showing a (p,n) process on target nuclei of masses of about 50–150 and for proton en-
ergies below 25 MeV. To illustrate the general principle, this graph only shows the (p,
n) exit-channel. Figure 13.14 reveals that after a maximum cross-section value at a cer-
tain energy, the cross-section decreases with further increasing projectile energy. Why
is this? With increasing energy of the projectile deposited in the compound nucleus,
another exit-channel “opens” because the higher excitation energy of the compound
nucleus allows more nucleons to “evaporate”. These alternative channels define new
394 13 Nuclear reactions
nuclear reactions in terms of A(a,b)B and have their own threshold values. Conse-
quently, exit-channels such as (p,2n) (not shown in Fig. 13.14) compete more and more
(with increasing projectile energy) with the first one, thereby lowering the probability
of the (p,n) channel.
1000
100
(mbarn)
Ě max
10
1
0 5 10 15 20 25 30 35
Fig. 13.14: Typical excitation function for a (p,n) nuclear reaction. Note the logarithmic scale of the
cross-section. The function indicates the threshold energy Ě of about 2.5 MeV. The nuclear reaction
starts at Eprojectile > Ě. From Ě onwards, the cross-section values increase with increasing projectile
energy. Here, the value of σ is about 10 mb at about 5 MeV. It further increases, but tends to
increase more slowly at about Eprojectile > 10 MeV. It then reaches a maximum value, here of about
500 mb at about 12 MeV. From now on, cross-section values decrease with increasing projectile
energy due to the onset of a “competing” exit-channel.
Figures 13.15 and 13.16 give concrete results for medically relevant iodine radioiso-
topes in comparison with other general diagrams. They show the experimentally
determined cross-section values for nuclear reactions leading to 124I, a positron
emitter used in molecular imaging. Figure 13.15 gives the 125Te(p,xn) reactions with
x = 1, 2 and 3, corresponding to product nuclei 125I, 124I and 123I. Figure 13.16 illus-
trates α-induced processes on 123Sb in terms of (α,xn) reactions with x = 1, 2 and 3,
corresponding to 126I, 125I and 124I, respectively.
13.4 Yields of nuclear reaction products 395
1000
n (mb)
cross section
10
0 10 20 30 40 50 60 70 80 90 100 110
proton energy (MeV)
Fig. 13.15: Excitation function of proton-induced nuclear reactions leading to iodine radioisotopes.
125
Te(p,xn) reactions with x = 1, 2 and 3 correspond to product iodine isotopes 125I, 124I and 123I.
With permission: A Hohn, FM Nortier, B Scholten, TN van der Walt, HH Coenen, SM Qaim, Excitation
functions of 125Te(p,xn)-reactions from their respective thresholds up to 100 MeV with special
reference to the production of 124I, Appl. Radiat. Isot. 55 (2001) 149.
1000
crosss section (mb))
10
1
5 10 15 20 25 30 35 40 45
-particle energy (MeV)
Fig. 13.16: Excitation function of α-induced nuclear reactions on 123Sb leading to iodine
radioisotopes. 123Sb (α,xn) reactions with x = 1, 2 and 3 correspond to 126I, 125I and 124I,
respectively. With permission: MS Uddin, A Hermanne, S Sudár, MN Aslam, B Scholten, HH
Coenen, SM Qaim, Excitation functions of α-particle induced reactions on enriched 123Sb and natSb
for production of 124I, Appl. Radiat Isot 69 (2001) 699.
396 13 Nuclear reactions
dŃB
R= = σΦa ŃA − λB ŃB (13:18)
dt
irr
ÁB = λB ŃB = σΦa ŃA 1 − e − λB t (13:19)
For example, 211At is produced via the 209Bi (α,2n) reaction irradiating metallic bismuth (see Vol-
ume II, Chapter 8), 124I through (p,n) or (d,2n) reaction on 124Te-TeO2, 18F through 18O(p,n) reaction
irradiating 18O-water, and 123I through p-induced reactions on 124Xe gas.
13.4 Yields of nuclear reaction products 397
Eo
Eprojectile
0
1 2 3 4 5 6 7 8 9 10
distance x
Fig. 13.17: Loss of kinetic energy of an incoming projectile when penetrating a target material along
a distance x. The real target dimension here is divided into 10 slices. These may be virtual, but may
also be realized practically, for example in metal targets by combining 10 thin metal foils. For an
overall 0.10 mm thick metal target, this would correspond to 10 separate foils of 10 μm thickness.
The projectile enters the first slice at its initial energy Eo. It then starts to lose kinetic energy in
terms of E = f(x). Within the first slices, the gradual loss of ΔE is low, but increases with decreasing
kinetic energy. It completely loses its kinetic energy at slice 9, and the final slice of the complete
target does not even “see” any projectile – and thus does not contribute to the nuclear reaction.
The same may be true for slice no. 9, if the kinetic energy of the projectile is below the nuclear
reaction threshold.
For a given target material, projectile penetration depends on the properties of the
projectile (what kind of projectile, how much kinetic energy). At the same time, a
given projectile shows very different penetration performances for different kinds of
Indeed, the production of 18F utilizes water enriched with >98% 18O. Now, H is 0.98 instead of
0.00205 and the yield with identical irradiations is about 500 times higher compared to irradiating
the same amount of natural water.
398 13 Nuclear reactions
10⁷ 10⁰
C Al C
Sn Al
10⁶ Fe
Ta Fe
U
Sn
Ta
10³
10²
10¹ 10¯²
10⁰
10¯¹
10¯² 10¯³
10¯² 10¯¹ 10⁰ 10¹ 10² 10³ 10¯² 10¯¹ 10⁰ 10¹ 10²
Proton energy (MeV) Proton energy (MeV)
Fig. 13.18: Range (a) and stopping power (b) of accelerated protons of increasing initial kinetic
energies for target nuclei covering light, medium and heavy chemical elements. C (Z = 6),
Al (Z = 13), Fe (Z = 26), Sn (Z = 50), Ta (Z = 73) and U (Z = 92). With permission: JF Ziegler, Transport
of ions in matter, International Business Machines, 1992.
Stopping power is given in units of MeV/cm. A more appropriate unit is the mass stopping
power, which is –dE / dx / ρ, in units of MeV g/cm2. In more detail, the stopping power includes
the effect of the projectiles interacting with both the electron shell of the target atoms (Selectric) and
with the nuclei of the target atoms (Snuclear).
13.4 Yields of nuclear reaction products 399
the projectile, εo the vacuum permittivity, e and me the charge and rest mass of the
electron.
2
dE Z
− =f 2 ρB (13:20a)
dx v
" 2 #
dE Ń A z 2 e2 2me c2 vc v2
− =Z ρ ln 2 − (13:20b)
dx M v 2 4πε 0 Upot ð1 − v c
c c
Figure 13.18 shows the range of accelerated protons of initial kinetic energies of
10 keV to 2 GeV for 6 target species covering light, medium and heavy chemical ele-
ments: carbon (Z = 6), aluminum (Z = 13), iron (Z = 26), tin (Z = 50), tantalum (Z = 73)
and uranium (Z = 92). Ranges continuously increase with increasing projectile energy
and are larger for low-Z targets relative to high-Z targets. The figure also translates
these ranges and energies into the stopping power for protons of kinetic energies up
to 100 MeV. Here, stopping powers increase from 10 keV to 100 keV, reaching maxi-
mum values. With further increasing kinetic energy of the protons, the values start to
decrease continuously.
Eprojectile
Eo
Eprojectile
= f(x)
= f(E)
0
1 2 3 4 5 6 7 8 9 10
distance x
Fig. 13.19: Excitation function σ = f(Eprojectile) with Eprojectile = f(x) of Fig. 13.17. Note that the
excitation function is left/right image reversed: the cross-section is zero within slice no. 8, i.e.
below the threshold energy, increases with increasing energy towards the surface of the target,
reaches a maximum value between slices 5 to 6 and decreases with further increasing energy
towards slice no. 1.
dΦ
− = σ Φ ŃA (13:22a)
dx
Φ = Φ 0 e − σ ŃA x (13:22b)
For neutrons, the interaction is much lower than in the case of charged projectiles.
For charged particles the so-called BRAGG effect can occur. This refers to an extremely
increased intensity of interactions at close to zero kinetic energy of the projectile. The
yield is an increased deposition of ionization within a very short distance, referred to
as the BRAGG peak.
13.4 Yields of nuclear reaction products 401
natRb(p,xn)82Sr natGa(p,xn)68Ge
distance x
Eo
E = f(x)
= f(E)
Eprojectile
100 90 80 70 60 50 40 30 20 10 0
Eprojectile (MeV)
Fig. 13.20: Schematic illustration of radionuclide production at Los Alamos National Laboratory for
(p,xn) reactions and proton energies of Eo = 100 MeV. The high proton flux density of 250 mA
makes the linear accelerator-based Isotope Production Facility a powerful radionuclide production
unit. Three different targets may be inserted in a row (stack), namely two separate natural rubidium
chloride targets and one natural gallium target. The protons first traverse the two Rb targets and
then arrive at the Ga target with less kinetic energy; i.e. ca. 100–70 MeV for the first (“high-
energy”) RbCl target, 60–40 MeV for the next (“low-energy”) RbCl target, and ca. 30–10 MeV for
the Ga target. These intervals each correspond to the maximum cross-sections of the three
different nuclear reaction routes, which are 87Rb(p,6n)82Sr, 85Rb(p,4n)82Sr and 69Ga(p,2n)68Ge.
After irradiation, the target stack is dismantled mechanically and each target is further treated
radiochemically to isolate the corresponding product nuclide, ultimately to extract 82Sr and 68Ge.
Adapted from: M Nortier, Recent advances in large scale isotope production at LANL, NRC8, Como,
September 2012. With permission.
o
o x
projectile
0 1 2 3 4 5 6 7 8 9 10
dx
distance x
Fig. 13.21: Decrease in projectile flux when penetrating a target of dimension dx.
402 13 Nuclear reactions
SF = 1 − e − ðt 2Þln2 = 1 − 0.5 − ðt 2Þ
irr =t1= irr =t1=
(13:23)
1.0
0.9
-(tirr/t½) ln2 -(tirr/t½)
SF = 1 - e = 1 - 0.5
0.8
0.7
satturation factor SF
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0 1 2 3 4 5 6 7 8 9 10
tirr / t½
Fig. 13.22: Correlation between irradiation periods and half-life of the product nuclide. Saturation
is achieved at SF = 1; however, this may need rather long irradiation periods depending on the
half-life of the nuclear reaction product. 50% of the saturation activity is achieved at an irradiation
period that equals one half-life of the product nuclide, and 75% at two half-lives etc.
In many cases, radionuclide production usually does not work under saturation conditions.
Most irradiations (in particular using charged particles, which need to be accelerated to high ki-
netic energies in high magnetic and electric fields) are quite expensive. Compared to an irradiation
lasting, for example, for two half-lives of the product nuclide (representing 75% of the maximum
achievable product activity) a further irradiation by one or two periods of half-life would add just
12.5% or 18.75% to the 75% value.
13.4 Yields of nuclear reaction products 403
When all the above-mentioned aspects are taken into consideration, the basic equa-
tion (13.18) turns into integral versions, shown in Fig. 13.23. For thermal neutrons
as projectiles eq. (13.19) applies, while for charged particles as projectiles eq. (13.25)
applies. The latter integral over the excitation function considers effects of thick tar-
gets, causing degradation in projectile energy and projectile flux.
FORMATION TRANSFORMATION
dŃB
R = ___ =
(114.21)
. NA . – . ŃB
a B
dt
. irr
Bt
(114.22)
YB = . ŃB = . NA . . (1 - e )
B a
H.NA
NA = ____
M
E2
H.NA dEa -1
____
____
(14.27a)
.tirr
YB = a
. . (1 - e B
) [ ] (E
( a) dE
d a
M d( .x)
E1
Fig. 13.23: Activation equations for neutrons (13.19) and charged projectiles (13.25). R is the
nuclear reaction rate, σ the cross-section for a certain exit-channel in barn (b), i.e. 10−24 cm2.
NA = NAVOGADRO. The projectile flux ϕa is cm−2 s−1 for neutrons and s−1 for charged projectiles.
In the latter case, it is typically expressed in units of Ampere (such as, e.g., μA).
For neutrons, the flux density is given in cm−2 s−1, which finally yields the unit for R
and ÁB (s−1 = Bq). For charged projectiles, the flux is given in projectiles per second
and, as for the unit known for electrons (the electricity), the Ampere is preferred.24
M is the target atomic mass (g), σ(Ea) is the excitation function with cross-sections
(mb), and S(Ea) is the stopping power in g/cm2 for 1 MeV. For protons, and for satu-
rated reaction yield, a specific expression is given in eq. (13.26). The product radio-
nuclide yield ÁB then has the unit Bq / μA.
The flux density of neutron is given in this unit. For charged particles (in analogy to electric
current) it is given in Ampere, typically μA or mA (1 A = 6.24⋅1018 electrons per second, 1 A = 1 C/s
or 1 A = 1 W/V).
404 13 Nuclear reactions
Eð2
H NAVOGADRO σ ð Ea Þ
Á B = 6.24 · 10 m
12
dEa (13:26)
M SðEa Þ
E1
Suppose a real target is (either virtually or physically) divided into many extremely
thin slices (dx ! 0). The nuclear reaction product B may be different and its prod-
uct radioactivity may be different in each slice according to the individual values of
the thickness of the slice, according to the projectile energy (a higher neutron en-
ergy may result e.g. in a (n,2n) nuclear reaction instead of (n,γ)), the corresponding
value of the cross-section, and the projectile flux. If the production yield is thus a
function of ÁB = f(dx) ≈ f(E), this can be measured experimentally by e.g. irradiating
a stack of several very thin samples of the target material. Suppose the individual
projectile energy within each sample is known, and the product activity is mea-
sured for each slice. The differential radionuclide production rate follows the exci-
tation function, i.e. it is low at low projectile energies, reaches a maximum at
higher projectile energy and drops off as projectile energy continues to increase.
In real radionuclide production, a thick target sample is irradiated rather than
several thin samples. The yield within the thick sample is thus cumulative (over
many virtual infinitesimally thin targets). It represents the integral of the individual
differential yields within a region of ΔEprojectile, i.e. between a starting (high) value
E1 and a lower one E2, as per Fig. 13.24.
Neutrons are very commonly used projectiles for two reasons: Neutrons do not need
to overcome the coulomb barriers, and they are available from various sources.
First, let us explore where the neutrons we need may origin. Table 13.3 lists various
sources of neutrons based on permanent facilities and laboratory samples. This repre-
sents a quite versatile and well-accessible spectrum of neutrons for many kinds of
applications. It includes nuclear reactions induced by charged particles of type (pro-
jectile,xn) to deliver high-energy neutrons, which in turn can be used to induce new
neutron-induced reactions. The neutrons “generated” show different parameters in
terms of energy and flux density. High initial kinetic energies can be moderated
down to lower values.
13.5 Radionuclide production using neutrons 405
E2 E1
1000
Y1
Y2
= f(Eprojectile )
10
0 5 10 15 20 25 30 35
Fig. 13.24: Differential vs. integral yields of a charged projectile-induced nuclear reaction. For a
given cross-section value at a specific projectile energy, differential nuclear reaction yields are
obtained according to eq. (13.18) if infinitesimally thin targets are irradiated. If a thick target is
irradiated, production yields are cumulative from Ethreshold to increasing Eprojectile. Suppose the
initial energy E1 of the projectile is 17 MeV and covers the whole thick target until the threshold
value, an integral yield ÁB(1) is obtained (here ca. 800 MBq/μAh) for 17 MeV ! threshold. Suppose
the target is not as thick, and the projectile beam leaves the target at 7 MeV, the projectile energy
acting in the target is 17 ! 7 MeV. In this case, the theoretical production yield ÁB(2) between
7 MeV ! threshold (which is ca. 200 MBq/μAh in this example) is not included. The overall
production yield is ÁB = ÁB(1) − ÁB(2) = 800–200 = 600 MBq/μAh.
Tab. 13.3: Neutrons available at nuclear reactors and from other sources used in neutron-induced
nuclear reactions.
Figure 13.25 shows the typical options of neutron-induced processes, namely (n,γ),
2x(n,γ) and (n,2n) (both along the line of isotopes), (n,p) (along the line of isobars),
and (n,d) and (n,γ) + β−.
The most frequently used pathway is the (n,γ) route, which produces a neu-
tron-rich radioisotope K2 of the same chemical element compared to the stable tar-
get isotope K1.25 At very high neutron flux densities, a subsequent neutron capture
takes place (double neutron capture), producing the mass (A + 2) isotope of the tar-
get nucleus. In both cases, the radionuclide cannot be separated chemically from
the target isotope.26 The radioisotope remains incorporated within the excess of
mass of the stable target isotope.
If the (n,γ) nuclear reaction product K2 is not of practical interest as directly
formed via eq. (13.27), but rather its radioactive subsequent (n,γ) ! β−-transformation
product, then the process in eq. (13.28) applies instead. The (n,γ) produced nucleus is
neutron-rich and thus transforms via a β−-process into a nuclide K3 of (Z + 1) proton
This also represents the most commonly applied process in Neutron Activation Analysis – see
Chapter 3 of Volume II.
. . . unless SZILLARD–CHALMER type processes are utilized, see Chapter 7 in Volume II.
13.5 Radionuclide production using neutrons 407
PROJECTILE = NEUTRON
3
b–
1 2
TARGET COMPOSITE
NUCLEUS NUCLEUS
(n,d) (n,p)
(n,a)
Fig. 13.25: Potential neutron-induced nuclear reactions and the subsequent β−-transformation
processes of the reaction products within the Chart of Nuclides. Color-coding in blue, because the
composite nucleus is neutron-rich. For these products in particular, β−-transformation is an
inherent strategy to obtain the (n,γ) ! β− ! product [1].
number relative to the target nuclide K1. K3 represents a different chemical element
relative to the irradiated target, and can be separated radiochemically. The product is
thus free of the initial target nuclide.
Figure 13.26 illustrates the two pathways of producing 177gLu, a low-energy β−-
emitter systematically used in nuclear medicine therapy. For more information, see
also Chapter 14 of Volume II.
408 13 Nuclear reactions
(n, )
176Lu 177Lu
2.59 % (n, )
160 d 6 71 d
6.71
2000
176Yb 177Yb
12 7 %
12.7
1.9 h
3
Fig. 13.26: Two pathways of neutron capture producing 177gLu.27 176Lu(n,γ)177gLu as direct process
and the 176Yb(n,γ)177Yb ! β− ! 177gLu route. Due to t½(177Yb) < t½(177gLu), already about one day
after the end of irradiation the irradiated 176Yb target exclusively contains pure 177gLu, i.e. all the
intermediate 177Yb has transformed into 177gLu. The product lutetium radionuclide is effectively
separated from the ytterbium target 176Yb.
In this case of the (n,γ) process on isotopically enriched 176Lu, the cross-section is very high
(2000 barn), but the abundance of the isotope 176 in natural lutetium is low (2.59%). Consequently,
isotopically enriched 176Lu is used (96%). Irradiations use neutron flux densities of 1014 to 1015 neu-
tron/cm2 s, and irradiation periods last for a 1, 2 or weeks. Batch activities approach the TBq level.
Specific activities after EOB (end of bombardment) reach values of about 20–30 Ci/mg, which is
below the theoretical maximum value (100 Ci/mg). Simultaneously, the metastable isomer 177mLu is
formed (t½ = 160 days). At EOB, it represents up to 0.05% radioactivity relative to 176Lu.The other
route starts from isotopically enriched 176Yb. The primary reaction product is 177Yb with a half-life
of 1.9 h. Already within the course of the irradiation, it transforms to 177gLu. After irradiation, 177gLu
is radiochemically separated from the 176Yb target of about 100 mg mass. It thus principally exists
at ultimate specific activity. The β−-transformation of 176Yb to 177gLu does not populate 177mLu. Prac-
tical specific activities reach 80% of the theoretical value.
The longer the “slow” neutron is in contact with the target nucleus, the higher the probability
that “something” happens, i.e. a neutron gets trapped.
Neutron energies and velocities correlate by E ≈ 5.23 v2 with E in meV units and v in km/s. At
0 °C, it is 0.0353 eV (ca. 2.6 km/s); at room temperature, 0.025 eV (ca. 2 km/s).
13.5 Radionuclide production using neutrons 409
10²
10¹
10⁰
10¯² 10⁰ 10² 10⁴ 10⁶
Neutron energy E (eV)
Fig. 13.27: Course of cross-sections of neutron-induced nuclear reactions over the range of kinetic
energies of 1 meV to 1 MeV. Neutrons are grouped into thermal, epithermal and medium to fast
neutrons.
According to increasing energy and velocity, neutrons are grouped into ultra-cold (<0.2 meV,
<0.2 km/s), cold (<2 meV, <0.6 km/s), thermal (<0.1 eV, <4.4 km/s), epithermal (<1 eV, <14 km/s),
medium-fast (0.5 eV–10 keV, 10–44 km/s), fast (10 keV–20 MeV, 44–62,000 km/s), and relativistic
(>20 MeV, >62,000 km/s).
By definition, resonance is the specific response of a system (basically in natural sciences) to
specific external activation. It is typically described in terms of the oscillation of the system with
significantly enhanced amplitude in the case that the excitation occurred by a characteristic fre-
quency of an electromagnetic activation – the system shows a resonant frequency.
410 13 Nuclear reactions
There is a broad range of thermal neutron capture cross-sections among the chemi-
cal elements. Values range from practically zero (e.g. for 4He, where the double magic
nucleon configuration is not at all receptive to a further neutron), through values
below 1 (i.e. from 0.0001 to 1 barn) up to several thousand barns. Cross-sections for the
individual isotopes of a given element again cover a range of values.32 The values also
reflect a consequence of the shell model of the nucleus: neutron cross-sections are typi-
cally larger when a nucleus with odd neutron number captures a neutron, yielding a
new nucleus of even neutron number. This alternation between the magnitude of
cross-sections of unpaired and paired neutron nuclei nicely illustrates the impact of
the shell model of the nucleus: paired nucleons provide a gain in binding energy.
Within nuclear reactors, nuclei undergo neutron capture reactions. For fissile nuclei
(see Chapter 12), there is another option: neutron-induced fission.33 As introduced in
the context of spontaneous fission, the fission process leads to the splitting of the nu-
cleus into two primary fragments, which (due to coulomb repulsion) move away from
each other with high kinetic energy. The excess neutrons they emit are of high kinetic
energy, up to 10 MeV. Suppose these neutrons pass the so-called moderators; in doing
so, they lose kinetic energy due to elastic scattering with the moderator atoms.34 When
moderated down to 0.025 eV, such a neutron induces subsequent processes on fissile
nuclei, especially 235U. Because the fission of, for example, one 235U atom releases
more than two single neutrons, these (moderated) neutrons may induce the fission of
two other 235U nuclei, leading to the second generation of fission. This turns a single
event into a fission cascade, shown in Fig. 13.28. It is used to produce nuclear power –
either in nuclear weapons or in nuclear energy plants, utilizing 235U, 239Pu or 233U.
Although this process is mainly seen in the context of nuclear energy, it is also a
significant source of many artificial radionuclides used today in science, industry,
and medicine. When a 235U target is irradiated in a nuclear reactor, neutron-induced
fission creates the typical profile of binary asymmetric fission, a characteristic distri-
bution of primary and secondary fission fragments and final neutron-rich fission
products. Some of these isobars include radionuclides of real practical value e.g. in
nuclear medicine, especially 99Mo, 90Sr and others. The radiochemical separation of
For example, for the stable isotopes of gadolinium (Z = 64), the corresponding σ-values are
152
Gd = 700 b, 154Gd = 60 b, 155Gd = 61 000 b, 156Gd = 2.0 b, 158Gd = 2.3 b, 157Gd = 254 000 b, 158Gd
= 2.3 b, 160Gd = 1.5 b.
This deserves a dedicated explanation; see Chapter 11 in Volume II.
In order to maximize elastic scattering and minimize neutron absorption, low Z elements are
preferred: hydrogen or deuterium as components of normal or “heavy” water, but also graphite (C),
cadmium, boron and lithium.
13.5 Radionuclide production using neutrons 411
Fig. 13.28: Cascade scheme of neutron-induced fission. The fission cascade is used to produce
nuclear power. Today’s nuclear power plants use 235U.
412 13 Nuclear reactions
Neutron capture on nonfissile nuclei may produce new generations of fissile nu-
clides, such as 239Pu and 233U. For 239Pu, the pathway is illustrated in Fig. 13.29.35 In
terms of nuclear power, the so-called breeding process sets in at the neutron-rich re-
action product 239U. It transforms by the primary β−-process to 239Np (t½ = 2.355 d),
which continues the same way to 239Pu (t½ = 2.4⋅104 a). This long-lived “bred” nu-
clide may be separated chemically from used 235U fuel and represents another gener-
ation of a fissile nuclide.
239P
2.411.104 a
u 3
4
239Np
FISSION
2.355 d
3
23.5 min
FISSION
Fig. 13.29: Neutron capture breeding reactions on uranium isotopes. Fissile 235U undergoes
neutron-induced fission (1), while 238U undergoes neutron capture (2) to 239U (t½ = 23.5 min). The
neutron-rich 239U transforms by β−-emission along an A = 239 = constant (3), until the most stable
nuclide of that isobar is reached, 239Pu. Following radiochemical separation of this fissile
plutonium isotope, it is ready to be used as a new target for neutron-induced fission (4).
For 233U, the analogue path starts from 232Th: 232Th (t½ = 1.405⋅1010 a) ! (n,γ) ! 233Th (t½ =
22.3 min) ! (β−) ! 233Pa (t½ = 27.0 d) ! (β−) ! 233U (t½ = 1.592⋅105 a).
13.5 Radionuclide production using neutrons 413
Tab. 13.4: Radionuclides produced via neutron-induced nuclear reactions and directions of their
applications. (IVMI = in vitro molecular imaging, NAA = neutron activation analysis, CRA = chemical
reaction analytics, RIA = radioimmunoassay assay, EBT = external beam therapy, ERT =
endoradiotherapy, SPECT = single photon emission tomography).
listed. Here, target isotopes of high natural abundance are preferred. If this value is
low but the cross-section high, target isotopes are isotopically enriched. The table
also lists typical applications of the produced radionuclides.
Neutron capture reactions cover a certain domain of product nuclides relative to the
target nucleus and product nuclides are neutron-rich in almost all cases. Charged
particles, in contrast, open access to a different neighborhood of the irradiated tar-
get nucleus, shown in Figs. 13.4 and 13.13. In most cases of large-scale radionuclide
production, nuclei of hydrogen and helium are used as projectiles, namely the pro-
ton 1H, the deuteron 2H, 4He and 3He. Because these projectiles introduce one or
two protons to the target nucleus, most of the product nuclei are proton rich. Chem-
ically, they thus differ from the irradiated target, can be isolated and finally repre-
sent “no-carrier-added” states.
13.6 Radionuclide production using charged particles 415
Charged projectiles must overcome the coulomb barrier to reach the nucleus, and
thus should have sufficient kinetic energy. This requires a technology that can in-
crease the kinetic energy of charged projectiles.36
13.6.1.1 Ionization
The general approach is to first ionize hydrogen or helium gas to produce ions. This
is done through dissociation of one (H) or two (He) electrons from the correspond-
ing nucleus, creating positively charged ions. This results in either 1H+ (the proton
itself, p) or 4He2+ (the alpha particle) – the nuclei of the ionized atoms. Optionally,
H2 can be ionized in such a way that negatively charged ions are obtained (1H−).
13.6.1.2 Acceleration
Subsequently, the ions are accelerated in electromagnetic fields. There are several
technological concepts, such as electrostatic VAN DER GRAAF generators, linear ac-
celerators and different versions of cyclotrons.
Interestingly, in the case of neutrons, the opposite applies. The challenge was to moderate their
kinetic energy.
The first artificial nuclear reactions were initiated by accelerated protons. Nobel Prize in Physics
in 1951 to DOUGLAS COCKROFT and ERNEST SINTON WALTON “ . . . for their pioneering work on the
transmutation of atomic nuclei by artificially accelerated atomic particles”.
416 13 Nuclear reactions
RF oscillator
UHF ≈ 10–300 kV
Vacuum tube
Ion source
Fig. 13.30: Concept of linear acceleration of ions. Ions released from the ion source are accelerated
in a long vacuum chamber within potential gradients between separate drift tube segments. Each
time the charged projectile leaves one drift tube and is in-between two segments, the
radiofrequency between these two segment switches.
energy of the ions, shown in Fig. 13.30. This concept may be extended significantly,
from several meters up to a length of kilometers. Energies reach values of GeV.
13.6.1.4 Cyclotrons
To achieve high kinetic energies of charged projectiles without an endless extension
of linear accelerator facilities, the idea of accelerating ions within cyclic paths was
proposed; it was one of many brilliant ideas in nuclear sciences.38 The concept of
acceleration is similar to the LINAC, having an alternating radio HF-based electric
field organized within a gap between two chamber segments. The two chambers
look like a flat tin can cut into two halves. Because they have a shape like the letter
D, they are called “dees”, shown in Fig. 13.31. Ions are released from the ion source
within the center of the magnetic field, and accelerating ions are forced to move
along a spiral trajectory by a static magnetic field B provided by two electromag-
nets. Each time an ion of charge q passes the acceleration gap (and this appears
always at the same frequency time t), the switch of the HF is according to its cyclo-
tron resonance frequency f, as per eq. (13.29) with m being the (relativistic) mass of
the ion.39 At that frequency, the centripetal force and the magnetic LORENTZ force
are equal. After each half cycle, the increased kinetic energy of the ion results in an
ERNEST LAWRENCE and STANLEY LIVINGSTON first realized a prototype of such a device (1930 at the
now Lawrence Livermore National Laboratory). LAWRENCE received the 1939 Nobel Prize in Physics
for “ . . . for the invention and development of the cyclotron and for results obtained with it, espe-
cially with regard to artificial radioactive elements”.
For nonrelativistic situations, the rest mass of the ions can be used, and cyclotrons of constant
magnetic field work at constant frequencies. This changes if projectiles such as the proton reach
energies of 100 MeV and above. 500 MeV protons already move at an incredible 75% of the velocity
of light.
13.6 Radionuclide production using charged particles 417
increased radius of its path.40 Finally, the ions are extracted41 and directed towards
the target. Their final kinetic energy thus depends on the HF acceleration and the
magnetic field, but also on the radius of the cyclotron. The projectile flux intensity
ranges from a tenth of μA to several mA.
qB
f= (13:29)
2π m
Electromagnet (N)
Acceleration gap
Ion source
Acceleration Ion source
gap Path
of the particle
Dees
Divesting +
–
capacitor D-shaped
electrodes
UHF („Dees“)
Extracted Projectile
Electromagnet (S) projectile
Fig. 13.31: Concepts of cyclotron acceleration of ions. (left) Assembly of the vacuum chambers (two
dees with the acceleration gap in-between) embedded by the strong electromagnets. (right)
Schematic drawing of the path of the ions from the ion source to the beam extraction position.
There are several categories of cyclotrons, summarized in Table 3.5. They range from
huge facilities capable of accelerating all kinds of ions (from protons to large ions
(such as Ca) at energies reaching GeV per mass unit) down to dedicated, often so-
called “medical” cyclotrons. These units provide protons, deuterons, 4He and 3He
projectiles only, with maximum kinetic energies of about 20 MeV for p and d and
about 30 MeV for 4He and 3He, respectively. Recently, so-called Mini-Cyclotrons have
become available. They are even smaller and provide protons only with maximum
The increasing mass of the projectile would “delay” the arrival time of the projectile at the accel-
eration gap between the dees. This requires technological solutions. One option is to change the fre-
quency of the radio-HF accordingly, and these types of cyclotrons are called “synchro”-cyclotron.
The other option is to adapt the magnetic field to the increasing velocity of the projectile. Cyclotrons
with corresponding alternating field gradients may use a constant radio-HF and are therefore called
“isochronous”-cyclotron.
For positive ions, this is accomplished by a deviating magnet. For accelerated negative ions,
such as H-, this occurs by directing the ions through a thin carbon stripping foil. The foil catches
the two electrons and thus converts an H- ion into a proton, which immediately moves differently
from the anions.
418 13 Nuclear reactions
Tab. 13.5: Types of accelerators used for radionuclide production. Courtesy: M Qaim, Cyclotron
production of medical radionuclides, in: Handbook of Nuclear Chemistry, volume 4, A Vèrtes, S
Nagy, Z Klencsár, RG Lovas, F Rösch (eds.), second edition, Springer, Dordrecht, Heidelberg,
London, New York, 2011.
Fig. 13.32: Vertical components of a “medical” cyclotron. Photograph of the PETtrace 700
“medical” cyclotron (courtesy: GE, Uppsala, manufacturer). The cyclotron is arranged in a vertical
position, allowing easy access by opening the door that represents one copper-based
electromagnet. Acceleration of H- ions is organized by 4 dees. Maximum proton energies are
9.6 MeV and maximum proton flux is 100 μA. Proton extraction occurs at the 9 o’clock position on
the left half, by mounting a carbon foil into the outer radius of the proton path. The protons
created are directed towards target stations. There are several target stations mounted, left.
Copyright General Electric Company. Used with permission.
13.6 Radionuclide production using charged particles 419
In order to produce the highest yield of a certain radionuclide, several aspects must
be considered in detail.
1. The excitation function of the nuclear reaction, its defining threshold energies
and the optimum range of the particle energy.
A target bombarded by charged projectiles of 10 MeV energy and 0.1 mA flux intensity accumu-
lates a power of 10 MeV × 0.1 mA = 100 W. For an irradiation period of 10 h, this gives 100 W ×
10 h × 3600 s/h = 36 MW s. Because projectile beams are only cm2 in diameter, this corresponds to
a high power density per target area.
420 13 Nuclear reactions
21Na
2
7
16F 17F 18F 19 F 20 F
1
109.7 m 100 %
4
15O 16 O 17 O 18 O 19O
99.762 % 0.038 % 0.002 %
6 5
Fig. 13.33: Charged particle-induced nuclear reactions leading to 18F. Isotope production starts
from stable nuclei. Reactions 1–3 are used to generate 18F, and the squares give the corresponding
compound nuclei C✶. (1) 18O(p,n), (2) 20Ne(d,α), (3) 16O(3He,p). Theoretically, there are many more
options. The squares indicate the target isotopes: (4) 17O(d,n), (5) 16N(α,2n), (6) 15N(α,n).
Details on the 18F-production routes, the targets used and the implications for subsequent nu-
cleophilic or electrophilic 18F-labeling chemistry are discussed in Chapter 13 of Volume II.
13.6 Radionuclide production using charged particles 421
Tab. 13.6: Experimental parameters of several production routes leading to 124I. Courtesy: SM
Qaim, Cyclotron production of medical radionuclides, in: Handbook of Nuclear Chemistry, volume
4, A Vèrtes, S Nagy, Z Klencsár, RG Lovas, F Rösch (eds.), second edition, Springer, Dordrecht,
Heidelberg, London, New York, 2011.
Nuclear Suitable energy range Thick target yield of I [MBq Impurity [%]
reaction [MeV] µA−h−]
I
I
Te(p,n) ! . <.
Te(d,n) ! . – .
Te(p,n) ! . .
Tab. 13.7: Radionuclides produced via charged particle-induced nuclear reactions and their main
applications. Projectile energies give approximate ranges. PET = positron emission tomography,
SPECT = single photon emission tomography, ERT = endoradiotherapy.
10⁴
¹²³I
10³
¹²⁴I
¹²⁴I
E₂ E₁
Yield (Mbq/mAh)
10²
10¹ ¹²⁵I
10⁰
10¯¹
0 10 20 30 40
Proton energy (MeV)
Fig. 13.34: Integral yield of iodine radioisotopes 125I, 124I, 123I in MBq/μAh. This corresponds to the
cross-section values for the nuclear reaction 125Te(p,xn) reactions with x = 1, 2 and 3 shown in
Fig. 13.15. The optimum proton energy covers the range from E1 = 18 MeV to E2 = 10 MeV. This
achieves a compromise between production yield and product purity. With permission from: A Hohn,
FM Nortier, B Scholten, TN van der Walt, HH Coenen, SM Qaim, Appl. Radiat. Isot. 55 (2001) 149.
13.7 Outlook
There are specific examples of nucleon transfer nuclear reactions that are not rele-
vant for the man-made production of artificial radioisotopes, but are important
from a scientific point of view. In addition to induced nuclear fission as a source of
atomic energy, there are two extremely relevant nuclear processes: nuclear fusion
and the genesis of the chemical elements in the stars.
424 13 Nuclear reactions
Figure 13.35 illustrates the main parts of the nuclear fusion cycle. First, two protons
within the sun plasma fuse together. This process is exclusively possible in the
core44 of the sun. This is not nucleon transfer, but nucleon fusion. The process does
not simply fuse two protons, creating a helium nucleus without a neutron; instead,
the heavy hydrogen isotope deuterium is formed, meaning one proton is converted
to a neutron via simultaneous release of a positron and an electron neutrino. In a
subsequent step, this deuterium nucleus fuses with another proton to create a stable
isotope of the next chemical element: Helium-3. High energy photons are simulta-
neously released. When two 3He nuclei fuse, 4He is created and two excess protons are
released, eqs. (13.30a–c).
p + p ! 2H + β + + ν e (13:30a)
2
H + p ! 3 He + γ (13:30b)
3
He + 3 He ! 4 He + 2p (13:30c)
The energy released in the nuclear reactions can be calculated according to the
gain in mean nucleon binding energies according to the LDM and the WEIZSÄCKER
equation for each step. These nuclear fusion reactions are thus exothermic, and re-
lease energy. To generate energy using man-made fusion, several options are being
discussed. The most promising one (largest cross-section) at low energy is reaction
(13.31). It fuses deuterium and tritium, the so-called DT fuel cycle. The 4He product
nucleus has a kinetic energy of 3.5 MeV, the neutron is emitted at 14.1 MeV maxi-
mum kinetic energy, the Q-value is 17.588 MeV.
2
H + 3H ! 4
He + n (13:31)
Stellar (and supernova) fusion reactions start with the proton plasma in the core of
a star, slowly and steadily converting the initial reservoir of hydrogen nuclei into
helium nuclei. In the later stages of a star’s life, i.e. when the abundance of helium
is large enough (and that of hydrogen nuclei small enough), fusion of helium nuclei
begins.
The core of the sun is a sphere of <0.7 solar radius. The core of our sun exists at temperatures of
about 15.6⋅106 °C. The temperature rapidly drops towards the outer regions, reaching 0.1⋅106 °C
already in the connective zone, and “only” ca. 5500 °C at the photosphere. The huge density and
temperature of the core allow the two protons to overcome the coulomb barrier.
13.7 Outlook 425
p e
d
3He
4He
p
3He
d
p
e
Fig. 13.35: Nuclear fusion in the core of stars. Two protons fuse to form deuterium. Another proton
is added to the deuterium to form a Helium-3 nucleus. Finally, two 3He nuclei fuse to form 4He and
two “free” protons (which become part of the next fusion cycle).
This process requires the star to have high temperature (>1.4⋅109 °C), mass and
density: such stars are several times larger than our sun (red giants). Here, three
4
He nuclei fuse to 12C. The newly formed nuclei of increasing mass combine with
each other (12C + 12C = 24Mg and smaller product nuclei) or continue to fuse with
4
He nuclei or free nucleons to form nuclei of even more chemical elements; at tem-
peratures >6⋅108 °C, masses of >6 times the mass of our sun, and densities of
>2⋅108 kg/m3. According to the LDM, this process remains exothermic until the
maximum mean nucleon binding energy is achieved, which is at iron, Z = 56.
From nuclei of iron (Z = 26) onwards, fusion reactions are endothermic. Thus,
these reactions cannot explain the genesis of heavier chemical elements. Addition-
ally, such reactions do not contribute to the generation of energy in stars – in con-
trast, they need energy due to their endothermic character. Instead, nuclear element
formation proceeds via neutron capture reactions. Neutrons appear in the stars
mainly as a side product of two fusion reactions:
13
C + 4 He ! 16 O + n (13:32a)
22
Ne + 4 He ! 25 Mg + n (13:32b)
426 13 Nuclear reactions
With a source of neutrons available, iron nuclei may now capture a neutron. The
nuclear reaction product would be an (A + 1) isotope of the initial iron nucleus; in
the case where this product is unstable and undergoes β−-transformation, the next
chemical element (Z + 1) is created. This galactic element evolution is indeed re-
sponsible for forming many of the stable isotopes between iron and bismuth (basi-
cally through the so-called s-process) and elements beyond bismuth (exclusively
through the so-called r-process).
Fig. 13.36: Profile of the s-process of element genesis. 56Fe is the starting nuclide, and (n,γ)
reactions are followed by direct β−-transitions.
Fig. 13.37: Profile of the r-process of element genesis. 56Fe is shown here as the starting nuclide,
but r-processes may start from almost any stable nuclide. Multiple (n,γ) reactions occur prior to
systematic β−-transitions.
Nucleon absorption within the whole nucleus induces nuclear reactions according
to nucleon transfer. This is the domain of relatively low kinetic energy projectiles
(≤ 10 MeV per nucleon), resulting in so-called low energy nuclear reactions. For pro-
ton-induced nuclear reactions, for example, on average the mean nucleon binding
energy of 8–9 MeV must be delivered to the compound nucleus in order to achieve
a (p,n) nucleon transfer. To release two neutrons, approximately twice as much en-
ergy must be deposited by the incoming projectile to allow for a (p,2n) exit-channel.
(p,6n)
(p,9n)
(p,4n)
(p,2n)
(p,3n)
(p,7n)
(p,5n)
(p,n)
?
Fig. 13.38: Nucleon transfer vs. spallation reactions. Processes of type (p,xn) cover product nuclei
of the same proton number Z for (p,n) for x = 2, 3, 4, etc. However, this cannot be extrapolated to x
ca. 10 and above. At about Ep > 100 MeV, the nucleon transfer mechanism is replaced by spallation
processes.
In the case of high energy charged projectiles (ca. 20–250 MeV per nucleon and ki-
netic energies >100 MeV, even reaching TeV levels), the projectile transfers its energy
directly to the nucleons of the target atom. This induces an intra-nuclear cascade of
energy transfer between individual nucleons. Finally, the very excited nucleus splits
into several pieces of varying mass number, instead of reacting to form one particular
product nucleus B. This is called spallation. Figure 13.39 schematically illustrates the
process. It is accompanied by the release of many individual nucleons, protons and
(in particular) neutrons of high kinetic energy. Three major steps can be distin-
guished, and protons are used as prototype of the projectile:
1. Initial proton (= projectile nucleon) interacts with target nucleon at very high
kinetic energy. Because the cross-section is σ = f(1/v), there is almost no reac-
tion and the protons pass the target nucleus as if it were “transparent”.
2. The incoming proton hits a target nucleon and eject it from the nucleus. This
generates a highly excited large fragment of the former target nucleus.
3. The excited nucleus de-excites. This occurs either by dividing into two fragments sim-
ilar to the fission process, i.e. with mass distributions around Afragment ≈ 1=2 A ± i, or
by evaporating a number of particles such as neutrons, protons, deuterons or α-
particles. This leaves many fragments of larger mass, where Afragment > 2=3 A.
13.7 Outlook 429
Spallation is interesting from several practical points of view. First, it produces nu-
clear reaction products that are not accessible by the nucleon transfer reaction
pathway. Second, the large number of neutrons emitted at high kinetic energy
makes them a new projectile for subsequent nuclear reactions. In this context,
spallation targets are also called “proton-neutron converters”. Third, spallation of-
fers hope of transmuting long-lived nuclear waste into short-lived, less ecologically
toxic products.
Fig. 13.39: Schematic illustration of spallation processes. Several product nuclei of varying mass
are released, together with single nucleons.
430 13 Nuclear reactions
Like the processes of spontaneous nuclear transformation, which can promote elec-
trons to higher energy levels in the formed nuclide, individual nuclear states may
also be produced in nuclear reaction processes, as shown in Fig. 13.40. The corre-
sponding nuclear states of a given product nucleus B are again defined by a charac-
teristic set of quantum numbers, with an emphasis on the overall moment J. Either
high-spin or low-spin levels may be preferred.
If a particular excited level of the product nucleus is formed, de-excitation within
the nucleus B occurs according to the character of secondary transformations as dis-
cussed in Chapter 11. Typically, high-spin states turn into low-spin states.
scattering product
nuclide
meta-stable
state
total
nucleon transfer exit-channel i
product
absorption nuclide
spallation
ground
state
fission
In many cases, one of the excited nuclear levels populated within the nuclear reac-
tion may even constitute a metastable state. Figure 13.41 illustrates the distribution of
metastable and ground state isomers of two technetium isotopes, 99 and 94, accord-
ing to the individual cross-sections. Both isotopes were produced in proton-induced
(p,xn) nuclear reaction, namely 100Mo(p,2n)99m,gTc and 94Mo(p,n)94m,gTc. The half-
life and overall spin/parity differ between metastable and ground state nuclei: 99mTc
(t½ = 6.0 h, JΠ = ½−) versus 99gTc (66.0 h, 9/2+), 94mTc (0.883 h, 2+) versus 94gTc
(4.9 h, 7+). The formation of the two different nuclear states delivers specific ratios
between the metastable and the ground state of the product nucleus. It may depend
on the kinetic energy of the projectile. For example, take Tc-94: the formation of
94m
Tc is favored at low proton energy (at Ep = 7 MeV, the ratio of 94mTc:94gTc is ≈10:1
in terms of differential production yield), while at higher proton energies (>Ep =
18 MeV) both states are populated in a 1:1 ratio, as shown in Fig. 13.42.
13.7 Outlook 431
500
400 94mTc
b]
crross section [mb
300
100 94gTc
0
5 10 15 20 25
0.9
0.8
94mTc 94m+gTc 99mTc 99m+gTc
on ratio R
R= / R= /
0.7
eric cross sectio
0.6
0.5
94Tc 99Tc
0.4
isome
0.3
0.2
0.1
0
5 10 15 20 10
proton energy [MeV]
Fig. 13.42: Ratio of cross-sections for the 100Mo(p,2n)99m,gTc and 94Mo(p,n)94m,gTc reactions.
Formation of 94mTc is favored at low proton energy. At Ep = 7 MeV, the ratio σ94m / (σ94m + σ94g) is
about 0.3, which corresponds to a 94mTc:94gTc ratio of ≈10:1 in terms of differential production
yield. At higher proton energies (>Ep = 18 MeV) the value of σ94m / (σ94m + σ94g) is ca. 0.5, i.e. both
states are populated in a 1:1 ratio. Experimental data according to Fig. 13.41.
14 Appendix
Tab. 14.1: Atomic units and selected fundamental constants.
. MeV
Energy equivalent mu c . MeV
RYDBERG constant R H
. m−
https://doi.org/10.1515/9783110742725-014
434 14 Appendix
Tab. 14.2: Selected atomic and nuclear constants of electron, proton, neutron, deuteron and alpha
particle. Atomic masses from: The 2020 Atomic mass evaluation. M Wang et al., Chinese Physics
C Vol. 45, No. 3 (2021) 030003.
Electron (e)
. MeV
Proton (p)
. MeV
Neutron (n)
. MeV
Deuteron (d)
J kg m− Hz K eV u
J ( J)/c ( J)/hc = . ( J)/h ( J)/k ( J) = . ( J)/c
= . × − = . × − = . × = . ×
m− ( m−) hc ( m−) h/c ( m−) c ( m−) hc/k ( m−) hc ( m−) h/c
−
= . × = . × − = . = . × − = . × − = . × −
Tab. 14.4: Masses of nuclides (atomic mass), mean nucleon binding energies ĒB, and mass excess
values Δmexcess for selected nuclides. From: M WANG et al., The AME2020 atomic mass evaluation,
Chinese Physics C Vol. 45, No. 3 (2021) 030003.
n . .
H . .
C . .
A area
A Ampere
Á (radio)activity
ÁS specific (radio)activity
THEORY
Ás theoretical (maximum) specific (radio)activity
ASARGENT coefficients
AGEIGER/NUTTAL coefficients
S
AV mass number of a sphere of volume V
S
AS mass number of a sphere of surface S
a D−S,R
coefficient in DASGUPTA–SCHUBERT, REYES equation
aSARGENT coefficients
a GEIGER/NUTTAL
coefficients
a als ra E
radius of an ellipsoid
−
Å Angström ( Å = m = pm)
qs
α quadrupole stretching factor
α alpha particles
α COULOMB
coefficients
α SURFACE
coefficients
B retardation (Bremszahl)
B SARGENT
coefficients
BGEIGER/NUTTAL coefficients
b SARGENT
coefficients
b GEIGER/NUTTAL
coefficients
b barn
β LDM
coefficient of the WEIZSÄCKER equation
C coulomb
Ć count rate
✶
C compound nucleus, composite nucleus
c speed of light
cD−S,R
coefficient in DASGUPTA-SCHUBERT, REYES equation
c as rb E
radius of an ellipsoid
d range
∅ diameter
Δ difference
Δ LAPLACE operator ∇⋅∇, ∇ or Δ
E energy
E C
coulomb barrier
EC electron capture
RECOIL
EK recoil energy
Ěa threshold energy
Ê energy operator
Én eigenvalues
Ei electric multipoles
14 Appendix 443
e electron
eV electron volt
ε electron capture
ε DEFORMATION
degree of nuclei deformation, eccentricity
G GAMOW factor
g gram
γ photon
H (isotopic) abundance
H HAMILTON function
Ĥ HAMILTON operator
H hindrance factor
H eigenvalue
h PLANCK constant
444 14 Appendix
I
I intensity
IC internal conversion
J joule
K EQU
chemical reaction equilibrium constant
⊙
K excited nuclear state
m
K metastable nucleus
g
K ground state nucleus
K B
BOLTZMANN constant
k C
COULOMB’S constant
L length ,
λ wavelength
M molarity
M muliplicity ,
M molecules
MP multipole
Mi magnetic multipoles
m mass
mo rest mass
Δm excess
mass excess
−
μ muon
N neutron number
N magic
magic proton number
N AVOGADRO
AVOGADRO constant or LOSCHMIDT number ( mol = .⋅ atoms)
n neutron xxx
n exponent
νe electron neutrino
446 14 Appendix
νμ muon neutrino
ντ tau neutrino
P parity operator
p impulse
p proton
Π parity
Ψ wave function
Q quadrupole
S
Q quadrupole of a sphere
E
Q quadrupole of an ellipsoid of rotation
14 Appendix 447
q electric charge ,
RH RYDBERG constant
r radius
ro radius parameter
rmass
mass radius
ro charge
charge radius parameter
S
r radius of a sphere
rC
radius at coulomb barrier
ρ psf
phase space factor
ρ density
S stopping power
s spin
S
S surface of a sphere
E
S surface of an ellipsoid of rotation
sf spontaneous fission
448 14 Appendix
σ cross-section
t time
t½ half-life
τ− tau
U potential energy , ,
U harmonic potential
U pot
ionization energy of the target material
U pot
ionization potential
V volume
ν̅ wavenumber
x fissility parameter
u frequency
ζLDM
coefficient of the WEIZSÄCKER equation
Index
²H 71 γ-photons 359
²²²Rn 168, 385 γ-quanta 311
²²³Ra 342 γ-radiation 309
²³²Th 62, 77, 412 γ-rays 315
²³³Pa 412 γ-spectrum 322, 349
²³³Th 412 γ-spectrum of 18F 349
²¹³Bi 342 γ-spectrum of a positron emitter 350
²¹¹At 342 γ-spectrum of ¹³¹I 325
³H 222
³He 222 0
n 224
¹H 69 100
Mo 236
¹²C 71 103
Pd 422
¹²N 71 10
B 262, 382
¹²³I 342 10
B-phenylalanin 263
¹³C 380 10
C 224
¹³¹I 322, 335, 342 111
Cd 190
¹³¹Xe 322, 333, 335 111
In 367, 422
¹¹¹In 342 112
Cd 190
α- and β-transformation 244 113
Cd 224
α/n conversion 267 116
Cd 236
α-emission 147, 150–151, 154, 164, 166–167, 11
B 188
240, 261, 264, 281, 294 11
C 112–113, 115, 123, 202, 206, 210, 224, 230,
α-emitting nuclides 268 367, 421
α-induced nuclear reactions 406 123
I 230, 367, 394, 419–420, 422
α-particle 7, 40, 150–151, 251, 263, 271, 294, 123
Sb 394–395
370, 374, 382 124
I 202, 206, 230, 365, 367, 394, 397,
α-process 154 420–422
α-radiation 265 124
Xe 419
α-scattering 10 125
I 202, 230, 361–362, 367, 394, 420
α-source 8 125
Te 394, 420
β+- and/or EC process 159 128
Te 236
β+-particles 349 12
C 56, 65–66, 72, 78, 102, 115, 123, 266, 271,
β+-process 158, 199 374, 425
β+-transformation 203 130
Te 236
β-parable 164–165 131
I 112, 202, 227, 413
β-particles 186, 201 131
Xe 227
β-process 150–151, 154, 160, 168 133
I 224
β-Transformation 156, 158–159, 167, 200, 218, 136
Xe 236
239, 273, 301 137
Cs 202, 224, 230, 233, 413
β-transformation balances 193 13
C 102, 123
β–-emission 281 13
N 202, 206, 210, 230, 367, 375, 380, 382,
β− and β+-particles 147, 370 421
β−-process 157, 199, 412, 426 148
Gd 251–253
β−-transformation 159, 301, 406 149
Tb 251
γ-emission 150, 309 14
C 102, 112–113, 115, 122, 202, 208–209,
γ-energy 173 229–230, 248, 271–272, 377, 384, 413
γ-knife 233, 340 14
N 113, 208, 374, 384, 386
https://doi.org/10.1515/9783110742725-015
452 Index
14 209
O 224 Bi 62, 136, 164, 396, 426
150 210
Nd 236 Bi 139–140, 143, 226
150 210
Yb 273 Pb 139–140, 143
151 210
Lu 273 Po 246, 251
153 211
Sm 230, 342, 413 At 118, 245, 251, 262, 396, 423
15 212
N 82 Bi 245, 251, 261, 264
15 212
O 202, 206, 210, 230, 367, 421 Po 261–262, 264
160 213
Tb 413 Bi 143, 251
161 213
Tb 367 Po 112
165 214
Er 202, 230 Bi 139, 168
166 214
Dy 413 Pb 139, 168, 226, 243
166 214
Ho 413 Po 168, 374
169 215
Er 413 Rn 270
16 218
O 78, 82, 91, 98 At 139
175
Yb 413 2-(18F)fluoro-2-deoxy-glucose 364
176 218
Lu 408 Po 139, 168, 243, 385
176 220
Yb 408 Rn 255
177g 222
Lu 407–408 Ra 272
177 222
Lu 202, 230, 262, 342, 414 Rn 243, 263, 406
177 223
Yb 408 Ra 251
17 224
F 82 Ra 251, 255, 342
17 225
N 82 Ac 143, 251, 423
17 226
O 91, 374, 386 Ac 239
180g 226
Ta 333 Np 239
180m 226
Ta 174, 333 Pa 239
186 226
Re 230, 342, 414, 422 Ra 119, 139, 143, 168, 239–240, 243–244,
188
Re 230, 342, 414 251, 263, 265, 270, 374–375, 385, 406
18 226
B 159 Rn 239
18 226
C 159 Th 239
18 228
F 159, 202, 206, 210, 224, 230, 233, 350, Th 255
229
364, 367, 380, 384, 419–421 Th 167, 251
18 22
N 159 Na 202, 224
18 230
Ne 159 Th 168, 243
18 231
O 91, 159, 271, 350, 384, 397, 419 Pa 167
193m 232
Pt 367 Th 101, 112, 136–137, 251, 254, 270, 278,
195m
Pt 367 281, 297, 306, 373
197 232
Au 374 Th transformation chain 255
197 233
Pt 414 Pa 166
19 233
F 73 Th 166
1 233
H 60, 64 U 167, 251, 297, 410, 412
234g
1st family of elementary particles 184 Pa 333
201 234m
TI 423 Pa 333
201 234
Tl 202, 342, 367 U 62, 168
206 235
Pb 61, 136, 246 U 62, 101, 136–138, 167, 251, 278, 281, 297,
206
Tl 246 306, 373, 410
207 235
Bi 359 U fission process 406
207 235
Pb 61, 136, 359 U fuel 412
208 236m
Pb 61, 136, 164, 262, 272 U 333
208 236
Tl 261 U 306
Index 453
237 39
Np 136–137, 143, 167, 251, 406 Ca 391
238m 39
U 333 K 78, 102, 391
238 3
Pu 251 H 112, 159, 202, 209, 224, 230, 248, 413
238 3
U 57, 62, 64–65, 69, 71–73, 78, 101, 112, He 59, 65, 159, 425
3
136–139, 143, 169, 250–252, 256, 270, Li 60
278–281, 298, 300, 305, 373 3-photon annihilation 353
238 40
U chain 246 Ca 72, 78
238 40
U transformation chain 140 K 78, 101–103, 112, 170, 224, 391
239 41
Np 412 K 78, 102, 391
239 42
Pu 167, 251, 298, 306, 410, 412 Ca 78
239 42
Pu 412 K 391, 413
239 42
U 412 Sc 391
241 43
Am 167, 251, 265, 406 Ca 78
242 44
Am 298, 333 Ca 78
242 44
Cm 298 Sc 365
242 45
Fm 298 Sc 78
242g 46
Am 333 Ca 78
242m 48
Am 333 Ca 236
242 4
Pu 298 He 40, 56, 59, 65, 72, 78, 241, 425
243 4
Am 298 Li 60
244 53
Fm 298 Fe 224
245 55
Cm 251 Co 421
247 56
Bk 251 Fe 57, 78, 426–427
248 57
Cm 281, 298 Co 230, 338
248 57
Fm 281 Fe 337
24 58
Mg 336, 425 Co 320
24 58
Na 202, 413 Fe 320
24 5
Ne 271 He 59
252 5
Cf 251, 281, 289, 291, 298, 303, 405 Li 60
252 60
No 298 Co 202, 227–228, 230, 233, 413
253 60
Rf 112 Ni 227
254 61
Cf 298 Cu 421
254 62
Fm 298 Cu 367
256 62
Fm 298 Ni 57
258 62
Sg 298 Zn 422
263 64
Db 298 Cu 170, 421
26 66
Al 230, 233, 320, 336, 377 Ga 422
26 67
Mg 320 Cu 342, 421
27 67
Al 385 Ga 342, 367, 422
2 68
H 65, 69, 72 Ga 202, 206, 230, 365, 367
68
2-photon annihilation 353 Ge 202, 401, 422
30 68
P 375, 385 Ge/68Ga 365
31 69
S 224 Ga 401
32 6
P 202, 230, 413 He 59, 224, 259
33 6
P 224 Li 59, 72
35 72
S 202, 224, 230, 413 As 365
36 73
S 78 Se 206, 422
37 74
Cl 78 As 365
38 75
Ar 78 Br 422
454 Index
75
Se 413 absolute activity 122, 124
76
Br 422 absolute energies of the electrons 363
76
Ge 236 absolute mass 49, 145
7
Be 204 absolute radioactivity 109, 120–121, 335
7
He 59 absolute values of Qα 248
7
Li 59, 204, 263 abundance 62, 397
81g
Kr 335 acceleration 415
81g
Rb 335 acceleration gap 416
81m
Kr 335 actinide elements 297
81
mKr 342 activation energy 386
82
Br 237 activation equation 396, 403
82
Kr 237 activity 106
82
Rb 365, 367 activity coefficients 381
82
Se 236 actor 184, 193
82
Sr 401, 422 age of the earth 138
82
Sr/82Rb 365 air 347
85
Kr 224 alarm signal 265
85
Rb 401 alchemists 374
86
Y 367, 422 alkali elements 81, 104
87
Br 302 allowed 222, 226
87
Rb 401 allowed and forbidden transitions 319
87
Sr 81 Alpha particle mass 434
88
Sr 81 alumina columns 136
89
Sr 81, 202, 230, 413 aluminium 52, 375
89
Zr 206, 365, 367, 422 aluminum-30 51
8
B 202 Ampere 403
8
He 259 amplitude 84
90
Nb 206, 365, 422 analytical technologies 368
90
Sr 413 ancient age 101
90
Y 202, 230–231, 262 angular distribution 208, 385
90
Zr 231 angular frequency 314
93
Zr 224 annihilation 233, 311, 365
94g
Tc 430, 432 annihilation photons 370
94m
Tc 422, 430, 432 annihilation radiation 344
96
Mo 236 annihilation velocities 354
96
Nb 237 annihilations constants 352
96
Zr 236 anode 367
99g
Tc 174, 234, 430 anti particles 180
99
Mo 136, 202, 230, 233–234, 334, 411, 413 anti-electron 186
99
Mo/99mTc 136 anti-hydrogen 351
99
Mo/99mTc radionuclide generator 175 antileptons 181
99
Mo/99mTc radionuclide generator 233, 411 antimatter 186, 311, 346
99m
Tc 112, 136, 174, 205, 234, 334, 339, 342, ARISTOTLE 1, 178
367, 422, 430 arrangement of shells 86
99m
Tc radiopharmaceutical 136 artificially produced unstable nuclides 102
99
Ru 334 assays 230
99
Tc 224, 233 astatine 122
9
Be 59, 267, 374 astrophysics 230, 233, 336
(A) and (CK) electron energies 362 asymmetric fission 293
Index 455
fission 277, 392 ground state 89, 147–148, 227, 235, 260, 305,
fission barrier 295, 305–306 309, 312
fission fragments 278 ground-state 174
fission isomer 333 ground-state electron 29
fission mechanism 289 ground state levels 171
fission process 273, 428 groups 16
fission vs. breeding 412 gyromagnetic parameter 318
fissionable 306
fissioning 155 hafnium 357
flavor 182 HAHN 278
FLEROV 278 HAHN 333
fluorescence 363 half-life 101, 107, 109, 113–114, 116, 131, 138,
fluorescence yield 363–364, 367 202, 216, 218–219, 236, 252, 269, 296, 321
flux density 400 half-life adequate to the medical purpose 339
follow-up processes 149, 298 half-life of the product nuclide 402
forbidden 222, 226 half-lives of nuclides undergoing spontaneous
formation of the positronium 349, 352 fission 299
formation probability 352, 354 half-lives of secondary transformations 331
fragments 288 half-lives of spontaneous fission 296
francium 42, 122 halo nuclei 72, 259
free neutron 186, 189, 235 halogen 104, 118
free proton 188 HAMILTONIAN function 29
frequency 13, 84, 296, 314 HAMILTONIAN operator 28
frequency factor f 259 Hamiltonian operator 195
ft-values 218–219 Handbook of Nuclear Chemistry 68, 76
fusion of helium nuclei 424 hard X-rays 316
fusion of light elements 377 harmonic oscillator 27, 85–88
fusion reactions 425 harmonic potential 27
heavier chemical elements 154, 425
gain in energy ΔE 146 heavy element research 376
galactic distribution of 26Al 338 heavy water 410
galactic element evolution 426 HEISENBERG 195
gallium target 401 HEISENBERG uncertainty principle 29
GAMOV 68 helium 41, 60
GAMOV factor 261 helium atom 40
GAMOW 259 helium isotopes 59
GAMOW factor 260 HEVESEY 140
GEIGER / NUTTALL 252 hexadecupol 49
GELL-MANN 178, 182 hexokinase 364
generator systems 135 HIGGS 198
genesis of heavier elements 377 HIGGS boson 377
geometrical deformation 287 high probabilities 321
glucose 364 high-energy charged projectiles 428
gluons 178 high-energy neutrons 303
glycolysis rate 364 high-frequency potentials 415
gold 8, 42, 49 high-spin or low-spin levels 430
gold-198 67 hillsides 161
GOLDSTEIN 6 hindrance factors 269
graphite 410 hindrances 221
460 Index
hole within the electron shell 148 internal conversion electron spectra 330
hot nucleus 377 internal conversion electrons 307, 327, 330
HUND 18, 23 intra-nuclear cascade of energy transfer 428
HUND’s rule 34, 75, 83 intrinsic spin 18
hydrogen 23, 25, 32, 36, 40, 42, 48, inverse process 235
59–60, 410 inverse β-transition 235
hydrogen atom 16, 351 iodine radioisotopes 394–395
hydrogen emission lines 13 ionisation of water 232
hydrogen emissions 6 ionisation potential 352
hydrogen spectrum 11 ionization 262, 265, 415
ionization energy 16, 80, 398
IC electrons 370 ionization of molecules 349
iIntegral yield of iodine radioisotopes 423 iron 65, 425–426
important β+- and AUGER electron emitting irradiation conditions 419
radionuclides 367 irradiation periods 402
important γ-emitting radionuclides 342 isobar 62–63, 98, 158, 199, 239, 243–244
impulse 29 isobar nuclides 100
impurity 421 isobar parabola 301
in-channels 379–380 isobars of integer mass number 163
independent particle model 94 isobars of integer mass numbers 165
individual coefficients 75 isobars of non-integer mass number 164
individual cross-section 392, 431 isochronous-cyclotron 417
individual differential yields 404 isodiaphere 63, 166, 241, 243
individual discrete energy 361 isomer 174
individual half-lives of excited levels 324 isomers in spontaneous fission 304–305
individual quantum number 67 isospin 58
individual secondary photon transitions 324 isotone 62
individual transition probabilities 217 isotones 63
individual vs. total cross section 392 isotope 41, 62–63, 158
induced fission 306, 405 isotope notation 62
inelastic scattering 349, 376 Isotope Production Facility 401
infinite square 87
infinite square potential 85, 88 JOLIOT 382
initial and final shell levels 356 JOLIOT F 375
initial and final states 379 Joule 39
initial and final wave functions 390
initial energy 397 KATRIN 213
initial states 194 kinetic energy 28, 154, 206, 251, 254, 258, 330,
inner bremsstrahlung 368–370 385, 408, 415
inner conversion 326 kinetic energy and range of the
integral radionuclide production yields 404 positron 346–347
intensity 322 kinetics 128
interaction of radiation with condensed kinetics of radioactive transformation 126
matter 368 kinetics of secular equilibrium 134
interactions with matter 369 kinetics of transient equilibrium 132
intermediate 126 K-shell 327, 354
intermediate unstable nuclides 136 K-shell electrons 205
internal conversion 326, 344 KURIE plot 214
internal conversion coefficients 328, 330 Kα and Kβ lines 359
Index 461
parent 137 photons 148, 173, 178, 235, 301, 309, 311, 370
parent nuclides 136, 139 photosphere 424
parity 58, 172, 220–221, 317, 430 PLÜCKER 4
parity inversion 221 PLANCK constant 13, 19, 314, 433
parity of electric multipole radiation 319 PLATO 1, 178
parity of magnetic multipole radiation 319 plum pudding model 7, 64
parity operator 221 plutonium 171
parity term 163 Pm 103
parity violation 221 point-like geometry 120
parsec 51 poison 117
partial nucleon binding energy 97, 104 polar coordinate system 31, 220
partial separation energy 81 polar variables 31
particle accelerators 415 POLIKANOV 304
particle emission 148, 271 polonium 7
particle in a box 84 polynomial 75
particle physics 376 position 25
particle research 377 positron 180, 233, 311, 330, 344–345
particle-wave duality 204 positron and its formation 346
partons 182 positron and molecular imaging 363
PASCHEN 13 positron annihilation processes 353
pathway of β-processes 157 positron emission 205
PAULI 189 Positron Emission Tomography 186, 353, 363
PAULI principle 23, 78, 83–84, 178, 182 positron-emitting radionuclides 186, 364
p-elements 19 positronium 349, 351–352
p-emission 147 post-division cascades 300
penetrability 259 post-effect 147, 341, 368
penetrability factor 259 post-processes 343
penetration 296 post-processes I 149
penetration length 398 post-processes II 149
Periodic Table of Elements 14, 61 potassium 391
periods 16 potential 85, 296, 387
pertechnetate 136 potential barriers 25
pertechnetate 205 potential energy 25–26, 28, 31
PET 206, 209, 233, 367, 420 potential energy of nucleons 85
PET-radiopharmaceutical 364 potential gradient 415
PETRZHAK 278 potential well 25, 84, 256, 387
PETtrace 700 418 precise mass of a nucleus 63
phase space factor 391 pre-formation 272
phase space volumes 194 primary fission fragment 279, 289, 296,
philosopher’s stone 374 298, 300
photoelectric effect 327 primary fission products 300
photon 233, 312 primary processes 307
photon emission 10, 307, 312, 334, 342 primary transformation 146–148, 241, 308
photon emission coefficients 324 primary β+-transformation 344, 349
photon energy 339 principal (or main) quantum number 18
photon multipole radiation 320 principle of PET detection 365
photon multipoles 321–322 probabilities for a neutron to induce a nuclear
photon transitions 320 reaction 390
photon-emitting radionuclide 339 probabilities of finding 28
Index 465