You are on page 1of 83

[Resumo] Fı́sica Nuclear e de Partı́culas

21 Sully
March 2023

TEM DOIS LIVROS DO MESMO CARA!! NUCLEAR AND


PARTICLE PHYSICS do Martin, e SÓ PARTICLE PHYSICS do
Martin TAMBÉM

Figure 1: Standard Model of Elementary Particles

1
Figure 2: Standard Model of Elementary Particles

1 Introduction
1.1 Relativity and Antiparticles
• Elementary particles physics = high-energy physics. Why high energy?

– E = mc2 ⇒ large energy needed


– To explore a particle, we need a probe particle with λ less than the
investigated particle. DeBroglie: λ = h/p ⇒ p must be large (FAST
particles man) ⇒ E must be large.
• Constructing a quantum theory that is consistent with special relativity,
we reach the conclusion that, for every particle of nature, there must be
an associated antiparticle.
• When brought together, particle-antiparticle pairs annihilate, releasing
their combined rest-energy 2mc2 as photons or other particles.

1.2 Symmetries and Conservation Laws


• Some symmetries lead to universal conservation laws
– translational symmetry ⇒ conservation of linear momentum
– rotational symmetry ⇒ conservation of angular momentum (has to
do with spins)

2
• Two symmetries of the Hamiltonian, which lead to important conservation
laws in nuclear/particle physics: parity and charge conjugation.

1.2.1 Parity
• Refers to spatial reflection, i.e. x → −x.
• P = ±1: intrinsic parity, or just parity.
• P is a multiplicative quantum number, that is conserved in strong and
EM interactions, but violated in the weak.
• P = 1 to all leptons and quarks, P = −1 to their antiparticles.

1.2.2 Charge Conjugation


• Refers to changing a particle to its antiparticle, i.e. x → x̄.
• Here, we will refer to states...
– without distinct antiparticles as a (e.g. photon γ)
– with distinct antiparticles as b (e.g. π + )
• Ca = ±1

1.3 Interactions
• Interactions involving elementary particles and/or hadrons are conveniently
summarised by ”equations”.
• Elastic scattering:
– Equations where LHS = RHS (same particles in the end but travelling
in different directions)
– Conserved quantum numbers must have the same total value in both
sides.
• ”Inelastic” scattering:
– In particle physics: equations where LHS ̸= RHS.
– In nuclear physics: when the final state is an excited state of the par-
ent nucleus A, that subsequently decays (e.g. via photon emission):
a + A → a + A∗ A∗ → A + γ

• Collisions between a given pair of initial particles do not always lead to


the same final state, but can lead to different final states with different
probabilities.
• Many particles are unstable and spontaneously decay to other, lighter par-
ticles.

3
1.4 Feynman Diagrams
1.4.1 Quantum Electrodynamics (QED)
• Primitive vertex: All electromagnetic phenomena are ultimately re-
ducible to the following elementary process:

Figure 3: Charged particle e enters, emits/absorbs photon γ, then exits. This


process is actually forbidden since it does not conserve energy, but it is a good
pedagogical example.

• To describe other processes we just patch primitive vertices together

Figure 4: Two electrons enter, a photon is emitted/absorbed (I need not say


which one emits the photon and which one absorbs it, the diagram repre-
sents both orderings), and the two then exit. In classical mechanics this is
the Coulomb repulsion of two equal charges. In QED it is Moller scattering.

• You can twist the diagrams around any topological configuration.


• If a particle is ”running backwards in time”, it is to be interpreted as an
antiparticle.
• The ”innards” of the diagram are irrelevant as far as the observed process
is concerned. Internal lines represent particles that cannot be observed
without entirely changing the process. We call them ”virtual” particles.
• Virtual particles can have any mass, energy is conserved in the end.
• Feynman diagrams do not represent particle trajectories. Horizontal
spacings do not correspond to physical separation.

• Each vertex in a diagram has a contribution to a total Feynman number


1
(don’t worry about it now), and introduces a factor of α = (e2 /h̄c) = 137 ,
the fine structure constant. Because it is small, new vertices contribute

4
Figure 5: Electron and positron annihilate to form a photon, which subsequently
forms an electron-positron pair. In classical mechanics this is the Coulomb
attraction. In QED it is Bhabha scattering.

less and less, such that you only need a few (≈≤ 4) to describe most
physical processes with great accuracy.

1.4.2 Quantum Chromodynamics (QCD)


• In QCD, colour plays the role of charge, and the fundamental process
(analogous to e− → e− + γ is quark → quark-plus-gluon (q → q + g)

• As before, we combine two or more of such ”primitive vertices” to represent


more complicated processes.

5
Figure 6: For example, the force between two quarks (which is directly respon-
sible for binding quarks together to make baryons - protons, neutrons etc. -
and indirectly responsible for holding neutrons and protons together to form a
nucleus) is described in this diagram.

• We say that the force between two quarks is ”mediated” by the exchange
of gluons.
• In the process q → q + g, the colour of the quark (though not its flavour)
may change. The gluon carries away the difference

Figure 7: A blue up-quark converting to a red up-quark. Since colour is always


conserved, this means that the gluon must carry away one unit of blueness and
minus one unit of redness.

• Gluons are thus ”bicoloured”, carrying one positive unit and one negative
unit.

• Since gluons themselves carry colour (unlike the photon, which is electri-
cally neutral), they couple directly to other gluons, and hence in addition
to the fundamental quark-gluon vertex, we also have two kinds of primitive
gluon-gluon vertices: three gluon vertices and 4 gluon vertices.

• In QCD, the coupling ”constant” α is not constant at all, it depends


on the separation distance between the interaction particles (we call it
a running coupling constant). It is big at large distances, and small at

6
short distances ⇒ assymptotic freedom (particles rattle around each other
instead of interacting).
• bunch of extra stuff in the book

1.4.3 Weak interactions (GWS)


• There is no name for the ”stuff” that produces weak forces (EM has
”charge”, strong has ”colour”). Some people call it weak charge.
• All quarks and leptons carry the weak charge.
• Two kinds of weak interaction:
– W: charged
– Z: neutral

LEPTONS-

• Fundamental charge vertex:

Figure 8: Negative lepton converts into corresponding neutrino, with emission


of a W − . Or, a positive lepton converts into corresponding antineutrino, with
absorption of W +

• We can combine vertices as always

Figure 9: µ− + νe → e− + νµ

• Fundamental neutral vertex:

7
Figure 10: In this case it works with any lepton (including neutrinos).

QUARKS-

• Fundamental charged vertex:

Figure 11: Quark with charge 1/3 converts into corresponding quark with charge
2/3 with the emission of a W − .

• The outgoing quark carries the same colour as the incoming one, but a dif-
ferent flavour (flavour = type of lepton/quarks). Flavour is not conserved
in weak interactions.
• Quark confinement ⇒ can’t see quarks like in the fundamental charge ⇒
quarks come with spectator quarks glued to it.

Figure 12: Beta decay of the neutron, i.e. n → p+ + e− + v e . The spectator


quarks are the u and d to the left.

8
• Quark Confinement: experimentally we know that quarks are only seen
in pairs of 2 (mesons) and 3 (baryons). As a consequence, the processes we
actually observe in the laboratory are necessarily indirect and complicated
manifestations of chromodynamics - it’s as if we could know electrodynam-
ics from the van der Waals forces between neutral molecules.

• Quarks are only allowed to move within their generation (u/d, s/c, t/b)...
except that in reality there is a bit of leeway for strangeness-changing
processes.

1.4.4 Weak and EM couplings of W and Z


Although this subsubsection is critical for the internal consistency of the theory,
it is of practical irrelevance for now.

• There are also couplings of W and Z to one another in GWS theory (just
as there are direct gluon-gluon couplings in QCD)

• Moreover, because the W is charged, it couples to the photon

1.5 Decays and Conservation Laws


• Every particle decays into lighter particles, unless prevented from doing so
by some conservation law. For example:
– the neutrinos and photons are stable because they have zero mass
and thus have nothing to decay into;
– the electron and positron are stable because they are the lightest
charged particles, so conservation of charge prevents their decay;
– the proton and antiproton are presumably stable because they are
the lightest baryons, so conservation of baryon number prevents their
decay;
• In practice, our world is populated mainly by protons, neutrons, electrons,
photons and neutrinos; more exotic things are created now and then (by
collisions), but they don’t last long.
• Each unstable species has a characteristic mean lifetime τ .

9
• Most particles exhibit several different decay modes. That means that an
unstable particle has different probabilities of decaying into different stuff.
• One of the goals of elementary particle theory is to calculate these lifetimes
and branching ratios.

• A given decay is governed by one of the fundamental forces (gravity is


dismissable, too weak). How can we tell?
– Photon comes out ⇒ the process is electromagnetic
– Neutrino comes out ⇒ the process is weak
– If neither comes out, well we are kinda fucked?

• Most dramatic experimental difference between strong, electromagnetic


and weak decays:
– Strong decay: ≈ 10−23 s
– EM decay: ≈ 10−16 s
– Weak decay: from 10−13 s to 15min (for the neutron)
• In general, things decay more rapidly the larger the mass difference be-
tween the original particle and the decay products, just as a ball rolls
faster down a steeper hill

• mp + me is almost mn . So, the decay n → p+ + e− + v e barely makes


it at all, and the lifetime of the neutron is by far greater than any other
unstable particle.
• Conservation Laws:

Figure 13: Since all physical processes are obtained by sticking these together
in elaborate combinations, anything that is conserved at each vertex must be
conserved for the reactions as a whole.

– Linear momentum and angular momentum: purely kinematic, con-


served in all interactions
– Charge: conserved in all interactions. In the case of the weak inter-
actions, the lepton or quark that comes out may not have the same
charge as the one that went in, but in that case the difference is
carried by W.

10
– Colour : conserved in strong interactions. The quark colour does
change, but the difference is carried away by the gluon (the direct
gluon-gluon also conserves colour). However, since naturally occur-
ring particles are always colourless, the observable manifestation of
colour conservation is pretty trivial: zero in, zero out.
– Baryon number : conserved in all interactions. In all the primitive
vertices, if a quark goes in, a quark comes out, so the total number
of quarks present is a constant. In this arithmetic we count quarks
as positive and antiquarks as negative. But we never see individual
quarks, only baryons (quark number: 3), antibaryons (quark number:
—3), and mesons (quark number: 0, they cancel out). So we just say
conservation of baryon number (A = 1 for baryons, A = — 1 for
antibaryons, and A = 0 for everything else). Since mesons carry
zero quark number, a collision can produce any number of mesons,
consistent with conservation of energy.
– Electron, muon and tau numbers: conserved in all interactions. If it
weren’t for Cabibbo mixing, there would be a similar conservation of
generation type for quarks (upness-plus-downness, strangeness-plus-
charm and beauty-plus-truth).
– Approximate conservation of flavour : conserved in EM and strong
interactions. It is not conserved in the weak force, but since that
force is so weak, we say that the various flavours are approximately
conserved. Some decays can only occur via the weak force, so some-
times flavour simply cannot be conserved.
– OZI rule: irrelevant for now

1.6 Particle Exchange


• Range of forces: Due to the uncertainty principle, the range of the
reaction A → A + X is

R=
MX c
– The EM force has infinite range because Mγ = 0
– The Strong force has infinite range because Mg = 0
– The Weak force has finite range since its exchange is with very heavy
particles (W ± and Z 0 )

• Range of the weak force: RW,Z ≈ 2 × 10−3


• Zero-range approximation: The range of the weak force is so small that
in many applications, we can consider it a zero-range or point interaction,
corresponding to the limit MX → ∞

11
1.7 Units and Dimensions
• Natural Units (nu): system of units used in particle physics. Chosen so
that the fundamental constants
– c = 1: unit of velocity
– h̄ = 1: unit of action (or angular momentum)
– eV : unit of energy
• Using these units we can, many times, just omit h̄ and c in equations since
they just multiply by 1.

• In natural units, all quantities have the dimensions of a power of energy.


In particular,
E
– M= c2
h̄c
– L= E

– T = E

Such that a quantity with meter-kilogram-second (mks) dimensions M p Lq T r


has the nu dimensions E p−q−r .

• Fermi Coupling Constant:

GF = 1.66 × 10−5 GeV −2

2 Leptonic Weak Interactions


• Coupling Constant relation:

GF g2 4παW
√ = W2 = 2
2 MW MW

12
Where gW = coupling constant associated with the W boson-lepton ver-
tices, and
g2
αW = W = 4.2 × 10−3 = 0.58α

Is the dimensionless strength parameter introduced by analogy with the
fine structure constant from QED’s α = e2 /4πϵ0 . From αW = 0.58α we
see that the EM and Weak forces are of comparable strength.

• Universality of Lepton Interactions: interactions between e− and νe are


the same as interactions between µ− and νµ and between τ − and ντ ,
provided the mass differences are taken into account.
• Decay Rate:
Γ(A → B) = KG2F MA5

• Branching Ratio B: Percent chance of this specific decay occuring in-


stead of the other possible decays.

• Lifetime of a particle:
B(A → B)
τl =
Γ(A → B)

• Neutrino Mixing:
This is the assumption that the neutrino states νe , νµ and ντ do not have
definite masses; instead they are linear combinations of three other states.
It is, however, a good approximation to only consider mixing between two
of the states, which we will call να and νβ . Then,

να = νi cos θij + νj sin θij

and
νβ = −νi sin θij + νj cos θij
Where θij = mixing angle.
• If θij = 0 then there’s no mixing.

3 Quarks and Hadrons


3.1 Quarks
bla bla bla nada que não tenha na tabela do inicio

13
3.2 General Properties of Hadrons
• Numbers for quarks:

– Up u: 1 para u, −1 para u
– Down d: 1 para d, −1 para d
– Charm C: 1 para c e −1 para c
– Strangeness S: −1 para s e 1 para s
– Top T : 1 para t e −1 para t
– Bottom B̃: −1 para b e 1 para b
• Other numbers:
– Baryon Number
1
B= (u + d + C − S + T − B̃)
3
FOR MESONS obviously B = 0
– Electric charge
2 1
Q= (u + C + T ) − (d − S − B̃)
3 3

• Each hadron is characterized by a set of values for B, Q, S, C, B̃ and T,


and in some cases there are no lighter hadron states with the same values
of these quantum numbers to which they can decay. These hadrons, which

14
cannot decay by strong interactions, are long-lived on a timescale of order
10−23 s and are often called stable particles. Here we shall call them long-
lived particles, because except for the proton they are not absolutely stable,
but decay by either the electromagnetic or weak interaction.
• The quark model says that in order for a particle to exist, its quark struc-
ture must be either 3q or qq
• You can also determine the quantum numbers by exploiting conservation
laws, i.e. if you know the type of interaction you can know which numbers
are conserved.

3.3 SÉRIE 1
1. Escreva sob a forma de equação os seguintes processos:
a) dispersão elástica de um antineutrino do eletrão com um positrão;
b) produção inelástica de um par de piões neutros em colisões protão-
protão;
c) aniquilação de um antiprotão com um neutrão para produzir três piões.
R:
a) Elástica: LHS = RHS ⇒ νe + e+ → νe + e+
b) p + p → π 0 + π 0 + p + p
c) p + n → π 0 + π 0 + π 0 ∨ π + + π − + π − . As duas configurações conservam
a carga (carga do antiprotão: −1. carga do neutrão: 0. logo a carga na
direita deve ser −1)
2. Desenhe um diagrama de Feynman de ordem mais baixa para os seguintes
processos:
3.

4.
5.
6.

7. Considere as sequintes combinações de números quânticos (Q, B, S, C,


B ), em que Q= carga elétrica, B = número bariónico, S = estranheza,
C= charme, B = “beleza”. Estes estados são compatı́veis com o modelo
de quarks?
• a) (-1, 1, -2, 0, -1):
2 1
Q= (u + C + T ) − (d − S − B̃) ⇔
3 3
2 1
⇔ −1 = (u + T ) − (d + 3)
3 3

15
e
1
B= (u + d + C − S + T − B̃) ⇔
3
1
⇔ 1 = (u + d + T + 3)
3
Resolvendo o sistema, ficamos com os números quânticos

u = u, d = 0, C = 0, S = −2, T = −u, B̃ = −1, B = 1, Q = −1

para ser aplicável ao modelo de quarks, tem de ter a forma 3q ou qq.


A única estrutura compatı́vel é então quando tomamos u = T = 0:

ssb

• b) (0,0,1,0,1):
Pelo mesmo raciocı́nio que a), chegamos nos números

u = u, d = 2, C = 0, S = 1, T = −u, B̃ = 1, B = 0, Q = 0

Como B = 0 sabemos ser um mesão, mas com d = 2, S = 1 e B̃ = 1


se torna impossı́vel encaixar em qq, pelo que é incompatı́vel com o
modelo de quarks.
8. Classifique cada processo como forte, fraco, electromagnético e dê as suas
razões.
• K + + p → Ω− + K + + K 0 : forte pois conserva todos os flavours e só
a forte faz isso
• Ω− → Ξ0 + π − : fraca pois flavour strangeness muda de 1 unidade
• Ξ0 → π 0 + Λ: fraca pois flavour strangeness muda de 1 unidade
• π 0 → γ + γ: EM pois há fotões
• K + → π + + π 0 : fraca pois flavour strangeness muda de 1 unidade
• µ+ → νµ : fraca pois há neutrinos
• K 0 → π + +π − +π 0 : fraca pois flavour strangeness muda de 1 unidade
9. Da seguinte lista de processos indique os que são permitidos ou proibidos
de acordo com as leis de conservação do Modelo Padrão da Fı́sica de
Partı́culas. No caso de ser permitido, indique qual interacção é a re-
sponsável, e no caso de ser proibido diga que lei de conservação é que
impede o processo.
R:
Temos que ver se Q, B, Le , Lµ e Lτ são conservados. Quando só há
hadrões, é fraca ou forte.
• a) p + p → π + + π 0
Impossı́vel, viola conservação de carga Q = 1 − 1 → 1 + 0

16
• b) η → γ + γ
Possı́vel (tudo foi conservado). Há γ ⇒ força EM
• c) Σ0 → Λ + π 0
Impossı́vel por conservação de energia, não sei pq
• d) Σ− → n + π −
Possı́vel (tudo foi conservado). fraca não sei pq (só hadrões=fraca
ou forte)
• e) e+ + e− → µ+ + µ−
Possı́vel (tudo foi conservado). EM não sei pq
• f) µ− → e− + νe
Impossı́vel, viola conservação de número leptônico muônico Lµ =
1→0+0
• g) ∆+ → p + π 0
Possı́vel (tudo foi conservado), forte não sei pq (só hadrões=fraca ou
forte)
• h) νe + p → n + e+
Possı́vel (tudo foi conservado), há neutrino (na esquerda) ⇒ fraca
(neutra Z)
• i) e− + p → νe + π 0
Impossı́vel, viola conservação de número bariônico B = 0 + 1 → 0 + 0
• j) p + p → Σ+ + n + K 0 + π + + π 0
Possı́vel (tudo foi conservado), forte não sei pq (só hadrões=fraca ou
forte)
• k) p → e+ + γ
Impossı́vel, viola conservação de número bariônico B = 1 → 0 + 0 e
leptônico eletrônico Le = 0 → −1 + 0.
• l) p + p → p + p + p + p
Possı́vel (tudo foi conservado), forte não sei pq (só hadrões=fraca ou
forte)
• m) n + n → π + + π − + π 0
Possı́vel (tudo foi conservado), forte não sei pq (só hadrões=fraca ou
forte)
• n) π + + n → π − + p
Impossı́vel, viola conservação de carga Q = 1 + 0 → −1 + 1
• o) K − → π − + π 0
Possı́vel (tudo foi conservado), fraca não sei pq (só hadrões=fraca ou
forte)
• p) Σ+ + n → Σ− + p
Impossı́vel, viola conservação de carga Q = 1 + 0 → −1 + 1
• q) Σ0 → Λ + γ
Possı́vel (tudo foi conservado). Há γ ⇒ força EM

17
• r) Ξ− → Λ + π −
Possı́vel (tudo foi conservado), fraca não sei pq (só hadrões=fraca ou
forte)
• s) Ξ0 → p + π −
Possı́vel (tudo foi conservado), fraca não sei pq (só hadrões=fraca ou
forte)
• t) π − + p → Λ + K 0
Possı́vel (tudo foi conservado), forte não sei pq (só hadrões=fraca ou
forte)
• u) π 0 → γ + γ
Possı́vel (tudo foi conservado). Há γ ⇒ força EM
• v) Σ− → n + e− + νe
Possı́vel (tudo foi conservado), há um neutrino ⇒ fraca

4 Symmetries
4.1 Revisão
• 1st order expansion: from the definition of a derivative,

f (x′ ) = f (x + ∆x) ≈ f (x) + f ′ (x)∆x, se ∆x << 1

• standard commutation relations: http://vergil.chemistry.gatech.edu/notes/quantrev/node18.html

4.2 Relationship between symmetries and conservation laws


(21)
1. Change of coordinates:
⃗ri → ⃗ri′

2. Hamiltonian Invariance:

H(⃗r1 , ⃗r2 , ...) = H(⃗r1′ , ⃗r2′ , ...)

3. Apply operator to arbitrary ψ

4. Apply operator to ψ ′ = Hψ
5. Substitute in observable

4.3 Translational Invariance ⇔ Cons. Linear Momentum


• Expresses the fact that all positions in space are physically indistinguish-
able.

18
• Change of coordinates:
ri → ri′ + δr

• Hamiltonian Invariance: For a single particle of mass m, the Hamiltonian


is  2
∂2 ∂2

1 ∂
H=− + 2+ 2
2m ∂x2 ∂y ∂z
It is invariant under the change of coordinates

H(r′ ) = H(r + δr) = H(r)

• Apply operator to arbitrary ψ: Here we introduce translation operator D̂

D̂ψ(r) = ψ(r + δr)

First order expansion:

D̂ψ(r) = ψ(r) + δr · ∇ψ(r)

Thus,
D̂ = 1 + iδr · p̂
Where p̂ = linear momentum opeator.

• Apply operator to ψ ′ = Hψ

D̂ψ ′ (r) = D̂H(r)ψ(r) ⇔

⇔ (D̂H(r) − H(r)D̂)ψ(r) = 0 ⇔
h i
⇔ D̂, H = 0

Substitute in observable: Substituting D̂, we find the conservation law for


linear momentum:
[p̂, H] = 0

4.4 Rotational Invariance ⇔ Cons. Angular Momentum


• Expresses the fact that all directions in space are physically indistinguish-
able.
• Change of coordinates: for example, rotation of angle θ around z axis

xi → x′i = xi cos θ − yi sin θ

yi → yi′ = xi sin θ + yi cos θ


zi → zi′ = zi

19
• Hamiltonian Invariance: if the system is rotationally invariant, the Hamil-
tonian will be the same. This holds for any closed system, as well as for a
particle moving in a central potential V (r) with the Hamiltonian
1 2
H=− ∇ + V (r)
2m
where
r = (x2 + y 2 + z 2 )1/2

• The derivation of angular momentum conservation from this invariance is


similar to the derivation of linear momentum conservation above, but is
more complicated. We will therefore give the derivation only for the case
of spinless particles, and state without proof the corresponding result for
particles with spin.
• Apply operator to arbitrary ψ: introduce R̂z operator

R̂z (δθ)ψ(r) = ψ(r′ ) = ψ(x − yδθ, xδθ + y, z)

First order expansion:


 
′ ∂ψ ∂ψ
ψ(r ) = ψ(r) − δθ y −x
∂x ∂y
Thus,
ψ(r′ ) = (1 + iδΘL̂z )ψ(r)
Where  
∂ ∂
L̂z = −i x −y
∂y ∂x
Is the z component of the orbital angular momentum operator, L̂. There
is, of course, nothing special about the z direction, and you may switch
that z for any direction n.
• Apply operator to ψ ′ = Hψ:

R̂n ψ ′ (r) = R̂n H(r)ψ(r) ⇔

⇔ ψ ′ (r′ ) = H(r′ )ψ(r′ ) = H(r)ψ(r′ ) = H(r)R̂n ψ(r) ⇔


 
⇔ R̂n H(r) − H(r)R̂n = 0 ⇔
h i
R̂n , H = 0

Substituting R̂n , we find the conservation law for angular momentum:


h i
L̂, H = 0

for a spinless particle whose Hamiltonian is invariant under rotation.

20
• If the particle has spin, then the total angular momentum J is the sum of
the orbital and spin angular momenta,
J =L+S
and the wave function Ψ may be written as the product of a space wave-
function ψ(r) and a spin wavefunction χ, i.e.
ψ(r)χ
For Spin- 21 particles:
– Spin-up: χ = α (Sz = + 12 )
– Spin-down: χ = β (Sz = − 12 )
• Pauli spin matrices σ:
1
Ŝ = σ
2
Where
     
0 1 0 −i 1 0
σx = , σy = , σz = ,
1 0 i 0 0 −1
Then α and β may be defined by
   
1 0
α= , β=
0 1
Under rotation, both spin and space wavefunctions will alter. Under a
rotation through δθ about direction n, the rotation operator generalizes
to
R̂n = (1 + iδθJˆ · n)
such that the transformation is
Ψ → Ψ′ = (1 + iδθJˆ · n)Ψ
L̂ and Ŝ act only on the space and spin wavefunctions respectively,
ˆ = (L̂ + Ŝ)Ψ = (L̂ψ(r))χ + ψ(r)(Ŝχ )

• We may use the same argument that led to conservation of angular mo-
mentum in the absence of spin may be applied to the rotation operator,
and lead directly to conservation of total angular momentum:
h i
ˆH =0
J,

• Rotation invariance does not lead to conservation of L̂ and Ŝ separately


and, in general, the orbital and spin angular momenta are not conserved:
h i h i
L̂, H = − Ŝ, H ̸= 0

because of the existence of spin-dependent forces.

21
4.4.1 Classification of particles
• ”particle” can be an elementary particle, or a composite particle, which is
a bound state of constituent particles.
• Spin: particle’s angular momentum in its own rest frame.

• The spin quantum number SP of a particle is the total angular momentum


J = L + S of its constituents. S is the spin of the constituents.
• SP = J can be either integer or half integer, while its z component can
take on any of the 2J + 1 = 2SP + 1 values.

• ”Good ” quantum number: constant of motion, commutes with Hamilto-


nian
• J is conserved, yet S and L are not separately conserved. Still, it is often
a great approximation to assume that L2 and S 2 are conserved. They are
defined as
L2 = L2x + L2y + L2z
and
S 2 = Sx2 + Sy2 + Sz2
Thus, h i h i
H, L̂2 = H, Ŝ 2 = 0

This conservation implies a symmetry, which means that, in this approxi-


mation, we can flip the spin, thereby changing its components yet not its
magnitude.

• In this approximation, L and S are also good quantum numbers, so that


the particle is characterised by SP = J, L and S, while a fourth Jz de-
pends on the orientation of its spin. This is the basis of the spectroscopic
notation.
• Spectroscopic notation:
2|S|+1
LJ
Where, instead of writing the numerical value of L, it is conventional to
write S, P , D, F , ... for L = 0, 1, 2, 3, ....

4.4.2 Angular Momentum in the quark model


• In the simple quark model, we make two assumptions:
– valid approximation to treat L and S as good quantum numbers
– the lightest states for any combinations of q q̄ (mesons) and 3q (baryons)
have L = 0 (S state).

• Mesons:

22
– q q̄ bound states
– single L in its frame
– q and q̄ are spin-1/2, so S = 0 ∨ S = 1
– if L = 0 ⇒ J = L + S = S ⇒ 2S+1 LJ = 1 S0 , 3 S1
– if L ≥ 1 ⇒ (S = 0 ⇒ J = L, S = 1 ⇒ J = L ± 1, L) ⇔

⇔ 2S+1 LJ = 1 LL , 3 LL+1 , 3 LL , 3 LL−1

– Here, J = SM = meson spin


• Baryons:

– 3q bound states
– two L in its frame: L12 and L3
– L12 : orbital angular momentum of a chosen pair of quarks about the
pair’s CM frame.
– L3 : orbital momentum of the third wheel about the CM of the pair
in the overall CM frame.
– Total orbital angular momentum:

L = L12 + L3

– Total spin:
S = S1 + S2 + S3
1 3
such that S = 2 ∨S = 2

23
– if L = 0 ⇒ 2 S1/2 , 4 S3/2
– if L = 1 ⇒ 2 P1/2 , 2 P3/2 , 4 P1/2 , 4 P3/2 , 4 P5/2
– if L ≥ 2 ⇒ 2 LL+1/2 , 2 LL−1/2 , 4 LL−3/2 , 4 LL−1/2 , 4 LL+1/2 , 4 LL+3/2

4.5 Parity ⇔ Cons. of Parity


• Change of coordinates:
⃗ri → ⃗ri′ = −⃗ri

• Hamiltonian Invariance: a system is invariant under parity if...

H(⃗r1′ , ⃗r2′ , ...) = H(−⃗r1 , −⃗r2 , ...) = H(⃗r1 , ⃗r2 , ...)

Parity is not conserved on weak interactions. For now, we will ignore the
weak interaction.
• Apply operator to arbitrary ψ:

P̂ ψ(⃗r, t) = Pa ψ(−⃗r, t)

Where
– a = particle type (e.g. e− , µ etc).
– Pa = constant phase factor.
Since two successive parity transformations leave the system unchanged,
we require that
P̂ 2 ψ(⃗r, t) = ψ(⃗r, t)
Implying
Pa = ±1
If we consider an eigenfunction of momentum

ψp⃗ (⃗r, t) = ei(⃗p·⃗r−Et)

Then
P̂ ψp⃗ (⃗r, t) = Pa ψp⃗ (−⃗r, t) = Pa ψ−⃗p (⃗r, t)
such that a particle at rest, with p⃗ = 0, is an eigenstate of the parity
operator with eigenvalue Pa .
For this reason, Pa = intrinsic parity, or parity, of particle a, with the
words ”at rest” left implicit.
• In addition to a particle at rest, a particle with a definite orbital angular
momentum is also an eigenstate of parity. The wavefunction for such a
particle has the form

ψnlm (r) = Rnl (r)Ylm (θ, ϕ)

24
Which leads to

P̂ ψnlm
⃗ (⃗
r, t) = Pa ψnlm r, t) = Pa (−1)l ψnlm
⃗ (−⃗ ⃗ (⃗
r, t)

and ψnlm (r) is an eigenstate of parity with an eigenvalue Pa (−1)l


• Apply operator to ψ ′ = Hψ: same thing
• Substitute in observable: same thing, thus we get
h i
P̂ , H = 0

4.5.1 Leptons and antileptons


• For all fermions (spin- 21 ):
Pf Pf¯ = −1
i.e. the fermion and antifermion have opposite parities
• They cannot be created or destroyed in EM interactions, thus we cannot
measure their intrinsic parities since nothing about them ever changes.
Hence we give them an arbitrarily defined parity:

Pe− := Pµ− := Pτ − := 1

Pe+ := Pµ+ := Pτ + := −1

4.5.2 Quarks and Hadrons


• Quarks can only created/destroyed as quark-antiquark pairs, so their in-
trinsic parities are also indeterminate and therefore arbitrarily defined:

Pu := Pd := Ps := Pc := Pb := Pt = 1

Pū := Pd¯ := Ps̄ := Pc̄ := Pb̄ := Pt̄ = −1

• Hadron = stuff made of quarks ⇒ we can know the parities of hadrons


from their quark structures.
– For a meson (M = ab̄),

PM = Pa Pb̄ (−1)L = (−1)L+1

– For a baryon (B = abc),

PB = Pa Pb Pc (−1)L12 (−1)L3 = (−1)L12 +L3

– For an antibaryon (B̄ = āb̄c̄),

PB̄ = Pā Pb̄ Pc̄ (−1)L12 (−1)L3 = −(−1)L12 +L3 = −PB

So that baryons and antibaryons have opposite parities.

25
• We could equally well have defined our conventions by fixing the parities
of five suitably chosen hadrons:

Pp := Pn := 1

and
PK − := PD− := PB − := −1

4.5.3 Parity of the charged pion


• We know the parity of π − from the reaction

π− + d → n + n

• Spin π − = Sπ− = 0, spin d = Sd = 1. J = L + S = 0 + 1 = 1.


• Deuteron = S-wave (L = 0) bound state of proton and neutron ⇒ Parity
= Pd = Pp Pn = 1
• Parity conservation: Pi = Pf ⇔ Pπ− Pd = Pn Pn (−1)L . So, Pπ− = (−1)L ,
where L is the orb. angular momentum of the final two-neutron state,
which must have J = 1 by angular momentum conservation.
• For reasons I didn’t understand L = 0, so

Pπ− = −1

4.5.4 Parity of the photon


• For reasons I didn’t understand,

Pγ = −1

4.6 Charge Conjugation


• ”Change of coordinates”: every particle becomes its antiparticle

a→a

• Conserved in EM interaction: Reverse particle = reverse charge and mag-


netic moment of every particle = EM interaction is invariant under charge
conjugation
• Conserved in strong interaction
• Not conserved in weak interaction, but we ignore the weak for now
• The Hamiltonian is invariant, bla bla bla and we have

[Ĉ, H] = 0 for strong and EM interactions

26
• a = particles that have a corresponding antiparticle
• α = particles that don’t have an antiparticle
• Under Ĉ we then have, for α particles:

Ĉ|αΨ⟩ = Cα |αΨ⟩

Thus, α particles are eigenstates of Ĉ with eigenvalues Cα = ±1, called


their C-parities.
For a particles:
Ĉ|aΨ⟩ = |aΨ⟩
We could have also put the Ca here but it would have no physical signif-
icance, as the relative phase of the two states |aΨ⟩ and |aΨ⟩ cannot be
measured.
• Since applying it twice reverts the antiparticle back to particle, we must
have
Ĉ 2 = 1 ⇔ Cα = ±1

• For many-particle states:

Ĉ|α1 , α2 , ...; a1 , a2 , ...; Ψ⟩ = Cα1 Cα2 ...|α1 , α2 , ...; a1 , a2 , ...; Ψ⟩

In the future, when the Ψ is not our primary concern, we shall omit it.
• For a particle-antiparticle pair, we have

Ĉ|aΨ1 , aΨ2 ⟩ = |aΨ1 , aΨ2 ⟩ = ±|aΨ1 , aΨ2 ⟩

and hence |aΨ1 , aΨ2 ⟩ is an eigenstate of Ĉ.


• For a fermion-antifermion pair, we have, from C conservation,

Ĉ|f f ; J, L, S⟩ = (−1)L+S |f f ; J, L, S⟩

• Quark model assumes that π 0 = 1 S0 state of uu or dd ⇒ S = 0, L =


0, J = 0 ⇒ the negative pion must have C-parity of

Cπ0 = 1

4.6.1 π 0 and η decays


• π 0 decays: experimentally confirms the quark model prediction Cπ0 = 1

π0 → γ + γ

• C conservation: Ĉ|γγ⟩ = Cγ Cγ |γγ⟩ = (−1)2 |γγ⟩ ⇔ Cf = Cγ2 = 1 ⇔ Ci =


Cπ0 = 1

27
4.7 SÉRIE 2
1. Utilizar as identidades dos comutadores para mostrar que, se

H = H0 + αL · S

tal que α = constante, e

[H0 , L] = [H0 , S] = 0

então L e S ainda são ”bons” números quânticos, i.e.

[H, S 2 ] = 0 [H, L2 ] = 0

R:

2. The deuteron is a bound state of two nucleons with spin-1 and positive
parity. Show that it may only exist in the 3S1 and 3D1 states of the np
system.
R:
J = 1 (spin of the composite particle, deuteron)
P = +1 (parity of the composite particle)
From parity conservation, Pi = Pf :

P = Pn Pp (−1)L = (−1)L = 1 ⇔ L = 2n : n ∈ Z

J =L+S ⇔L=J −S ⇔L=1−S


Sn = Sp = ±1/2, S = Sn + Sp ⇒ S = 0 ∨ S = ±1
S = −1 ⇒ L = 2 = 2n (ok)
S = 0 ⇒ L = 1 ̸= 2n (impossı́vel)
S = 1 ⇒ L = 0 = 2n (ok)
ou seja, só podemos ter

J = 1, L = 2, S = 1 J = 1, L = 0, S = 1
2S+1
LJ ⇒ 3 S1 , 3 D1

3. If pp annihilation at rest proceeds via S states, explain why the reaction


p + p → π 0 + π 0 is forbidden as a strong or electromagnetic interaction.
R:
”proceeds via S states” ⇒ L = 0
Parity conservation: Pi = P f except for weak interactions, i.e. it is con-
served for strong and EM.

Pi = Pp Pp (−1)L = 1 · (−1) · (−1)0 = −1

28
Pf = Pπ0 Pπ0 (−1)L = Pπ20 (−1)0 = 1 · 1 = 1
Thus
Pi ̸= Pf
Cannot occur via strong or EM.
4. The ground state baryons have L12 = L3 = 0 in the simple quark model.
If the next lowest-lying band of baryons has L12 = 1, L3 = 0 or L12 = 0,
L3 = 1, what values of the spin J and parity P occur?
R:
LB = L12 + L3 = 1 em ambos os casos.
Paridade: PB = Pq Pq Pq (−1)L12 +L3 = 1 · 1 · 1 · (−1)LB = −1 P
Bariões são formados de 3 quarks, cada um com spin ±1/2 ⇒ S =
±1/2 ∨ ±3/2 ⇒ S = 1/2 ∨ S = 3/2
J = L + S = L12 + L3 + S = 1 + S
S = 1/2 ⇒ J = 1/2 ∨ J = 3/2
S = 3/2 ⇒ J = 1/2 ∨ J = 5/2

5. Suppose that an intrinsic C-parity factor Ca is introduced into


Ĉ|a, Ψ⟩ = |a, Ψ⟩
which then becomes
Ĉ|a, Ψ⟩ = Ca |a, Ψ⟩
Show that the eigenvalue corresponding to any eigenstate of Ĉ is indepen-
dent of Ca , so that Ca cannot be measured.
R:
Como já foi visto, aplicar Ĉ duas vezes numa partı́cula faz p → p → p,
i.e. não muda nada. Logo, Ĉ 2 = 1 e
Ĉ 2 |a, Ψ⟩ = |a, Ψ⟩ (1)
Isto é, |a, Ψ⟩ Mas,
Ĉ 2 |a, Ψ⟩ = Ca Ĉ|a, Ψ⟩ = Ca Ca |a, Ψ⟩ (2)
Então, (1) = (2) e
Ca Ca |a, Ψ⟩ = |a, Ψ⟩ ⇔ Ca Ca = 1
e daı́??????????????
6. The η(549) meson has spin-0 and is observed to decay to three-pion fi-
nal states by the electromagnetic processes (η → π 0 + π 0 + π 0 ) and
(η → π + + π − + π 0 ). Use this information to deduce the parity of η,
and hence explain why the decays η → π − + π + and η → π 0 + π 0 have
never been observed.
R:

29
• Conservação de paridade:

Pi (= Pη ) = Pf ⇔ Pη = Pπ3 (−1)L12 (−1)L3 = (−1)L12 +L3 +1

• ”η(549) meson has spin-0” ⇒ J = 0

J = L + S = L12 + L3 + S

• S = 0 para as duas reações, pois Sπ0 = Sπ− = Sπ+ = 0:


• Assim, temos
J =L⇔L=0

• Portanto,
Pη = (−1)1 = −1
• Se tivéssemos as outras duas reações, η → π − + π + e η → π 0 + π 0 ,
o mesmo cálculo daria uma paridade diferente, Pη = 1. Como só se
observam as duas primeiras reações, sabemos que o valor verdadeiro
da paridade é Pη = −1
7. Considere a reacção π − + d → n + n, em que d é um estado ligado de um
protão e um neutrão (chamado deuterão), com spin 1, no estado S. O pião
inicial está em repouso. Deduza a paridade intrı́nseca do pião negativo.
R:

• ”d tem spin 1, estado S” ⇒ Sd = 1, Ld = 0


• ”o pião inicial está em repouso” ⇒ acho que daqui tiramos que Lπ− =
0 (esta condição é necessária para fazer a questão)
• Conservação de paridade:

Pi = Pf ⇔ Pπ− Pd = Pn2 (−1)L ⇒ Pπ− = (−1)L

• Conservação de momento angular:

Ji = L + S = Lπ− + Ld + Sπ− + Sd = 0 + 0 + 0 + 1 = 1 = Jf

Jf = 1 ⇔ 1 = L + S
Mas neutrões tem S = 0, portanto L = 1 e Pπ− = (−1)1 = −1

5 The Quark Model


5.1 Isospin symmetry
• Within a given family, all particles have same spin, parity, baryon number
and S, C, B̃, but differ in Q.

30
• Isospin Symmetry: good approximation to mu = md and the forces acting
on each being exactly equal.
• The strong forces on u and d quarks are the same, as far as we know,
but the EM forces are different because the quarks have different electric
charges. In addition, the d quark is a slightly heavier than the u quark,
as we shall show below. However, this quark mass difference is small
compared with typical hadron masses, and electromagnetic forces are weak
compared with strong forces. Consequently, isospin symmetry is a good
approximations. Families of particles like n, p (nucleons) and K + , K 0 , K −
(kaons) are called isospin multiplets.

5.1.1 Isospin Quantum Numbers


• Hypercharge Y : constant for members of same isospin multiplet.

Y = B + S + C + T + B̃

• As these quantum numbers have the same values for all members of a
given isospin multiplet, so does the hypercharge.

• I3 quantum number:
Y 1
I3 = Q − = (u − d)
2 2

• isospin I = Maximum value of I3 within a multiplet:

I = (I3 )max

• All observed multiplets contain, precisely, (2I +1) members, with I3 values
of
I, I − 1, I − 2, ..., −I

• Sum of two Isospins I a and I b can give the values:

I a + I b , I a + I b − 1, ..., |I a − I b |

• Sum of their third components I3a + I3b :

I3 = I3a + I3b

• Particle has I3 ̸= 0 ⇒ as a consequence of isospin symmetry, other parti-


cles belonging to the same isospin multiplet must exist.

31
5.2 Lightest Hadrons
• Lightest quarks: u, d, s.
• Lightest hadrons are composed of the lightest quarks, and have L = 0,
zero orbital angular momentum.

• In this subsection we only discuss these lightest particles, so here C =


B̃ = T = 0 ⇒ Y = B + S

5.2.1 Light Mesons


• Two families of light meson, each with nine-particles:
– pseudoscalar meson nonet: has spin-parity J p = 0−
– vector meson nonet: has spin-parity J p = 1−

5.2.2 Light Baryons


• Supermultiplets: families of hadrons with the same B, spin and parity, like
the 0− and 1− meson nonets.
• All hadrons belong to such families.

• For mesons, they are nonets; for baryons they can be singlets, octets or
decuplets (10), respectively.
• For simple systems, the lowest-lying states almost always have spatial
wavefunctions which are symmetric under the exchange of like particles,
and we will assume this here. Since we are assuming space-spin wavefunc-
tions, this implies that the spin wavefunction must also be symmetric un-
der exchange of like quarks; i.e. any like pair of quarks aa must have
spin-1, corresponding to parallel and symmetric quark spins,
instead of spin-0, corresponding to antiparallel, antissymetric
spins.

• Let’s consider all possible baryon quark structures using u, d, s and L = 0:


3
– aaa: uuu, ddd, sss ⇒ J = 0 + S = 2
– aab: uud, uus, ddu, dds, ssu, ssd ⇒ J = 0 + S = 12 , 32 .
– abc: uds ⇒ J = 0 + S = 21 , 12 , 32 . These arise because the ud pair can
have spin-0, leading to S = 1/2 on adding the spin of the s quark, or
spin-1, leading to either S = 1/2 or S = 3/2 on adding the s quark
spin
• Although we have considered just the three lightest quarks, it is straight-
forward to extend these ideas to include more quark flavours, although
the weight diagrams become increasingly complex.

32
5.3 SÉRIE 3
1. O hadrão Σ+ C (2455) apresenta o seguinte modo de decaimento: ΣC →
+
+ 0 +
ΛC + π , com uma taxa tı́pica das interações fortes, em que ΛC = udc é
um hadrão isosingleto (ver tabelas). Deduza Q, B, S, C, B̃, T de Σ+ C e, a
partir daı́, sua composição de quarks. Σ+
C possui parceiros de isospin? Se
sim, quais seus conteúdos de quarks?
R:

π 0 = uu ∨ dd ⇒ Σ+ +
C = ΛC = udc
1 1 1 1
I3 (π 0 ) = (u − d) = (0 − 0) = 0, I3 (Λ+
C) = (u − d) = (1 − 1) = 0
2 2 2 2
I( π 0 ) = 1 (tabela), I( Λ+
C ) = 0 (tabela), I(Σ+
C ) = 1 (tabela)

Logo
I( Σ+ + +
C ) = I( π ) + I( ΛC ) = 1 + 0 = 1

Logo há mais duas partı́culas na famı́lia do Σ+


C , uma com I3 = 1 e outra
com I3 = −1. Então, haja vista que Y = B + S + C + B̃ + T = 1 + 0 +
1 + 0 + 0 = 2,

I31 = Q − Y /2 ⇔ 1 = Q − 1 ⇔ Q = 2

I31 = Q − Y /2 ⇔ −1 = Q − 1 ⇔ Q = 0
Como o charm C é igual, sabemos que as partı́culas da familia devem ter
um, e apenas um, quark c. Além disso, S, B̃, T também são constantes
para a mesma familia de isospin, então sabemos que as partı́culas não
podem ter s, b, t. As únicas possibilidades para as cargas Q = 2 e Q = 0
são, respectivamente, uuc e ddc.

2. Estime a distância média atravessada por uma ressonância produzida


numa reação do tipo π − + p → X 0 + n, com γ = E/m ≈ 10, antes
que decaia através de um processo do tipo X 0 → π + + π − . Compare esta
com a distância mais pequena que pode ser resolvida experimentalmente
e com o alcançe da interação forte. Porque é que esta última comparação
é importante?
R:
Distância média: d = γcτ , ou seja, velocidade da luz × tempo médio de
vida × fator γ = E/m.
Considerando-se uma interação forte, τ ≈ 10−23 , pelo que teremos

d = (10)(3 × 108 )(10−23 ) = 3 × 10−14 m = 30f m

Agora falta descobrir o alcance da interação forte. Usando o princı́pio da


incerteza:

∆x∆p ≥
2

33
Mas p = E/c, então ∆p = ∆E/c:

∆x∆E h̄

c 2

Mas, pela outra versão do princı́pio da incerteza, ∆E∆t ≥ 2, ou seja

∆E ≥ 2∆t , e temos portanto

∆xh̄ h̄

2∆tc 2
Logo,
∆x =≥ ×10−15 m = 3f m
Isto é, o alcance da interação forte é da ordem de 10−15 m, sendo que
a distância atravessada é maior, sendo da ordem de 10−14 m. Esta com-
paração é importante porque nos mostra que X 0 não decairá pela interação
forte antes, tal que o decaimento π − + p → X 0 + n ocorre numa boa?
3. Mostre que um mesão que decai para pares π + π − através da interação
forte tem de ter C = P = (−1)J , em que J é o spin do mesão. Os
mesões ρ0 (770) e f 0 (1274) decaem através da interação forte para dar
pares π + π − e têm spin-1 e spin-2 respectivamente. Qual dos decaimentos
ρ0 → π 0 +γ e f 0 → π 0 +γ é proibido na interação electromagnética? Qual
dos decaimentos ρ0 → π 0 + π 0 e f 0 → π 0 + π 0 é proibido em qualquer
interação?
R:
Já que piões decaem para dois fotões (π → γ + γ), por conservação de
momento angular sabemos que eles têm de ter S = 0.

Sπ = 0 ⇒ J = L

P = Pπ2 (−1)L = (−1)J


C = (−1)L+S = (−1)J
A partir daqui não sei se está certo:
• ρ0 → π 0 + γ:
– Ci = Cf ⇔ (−1)J = Cπ0 Cγ ⇔ −1 = 1 × (−1) ⇒ OK
– Pi = Pf ⇔ (−1)J = Pπ0 Pγ ⇔ −1 = 1 × (−1) ⇒ OK
• f 0 → π 0 + γ:
– Ci = Cf ⇔ (−1)J = Cπ0 Cγ ⇔ +1 = 1 × (−1) ⇒ mal
– Pi = Pf ⇔ (−1)J = Pπ0 Pγ ⇔ +1 = 1 × (−1) ⇒ mal
• ρ0 → π 0 + π 0 :
– Ci = Cf ⇔ (−1)J = Cπ0 Cπ0 ⇔ −1 = 1 × 1 ⇒ mal
– Pi = Pf ⇔ (−1)J = Pπ0 Pπ0 ⇔ −1 = 1 × 1 ⇒ mal

34
• f 0 → π0 + π0 :
– Ci = Cf ⇔ (−1)J = Cπ0 Cπ0 ⇔ +1 = 1 × 1 ⇒ OK
– Pi = Pf ⇔ (−1)J = Pπ0 Pπ0 ⇔ +1 = 1 × 1 ⇒ OK
4. O π + e os seus estados excitados com massas abaixo de 1.5GeV 2 são
mostrados na figura ao lado. Identique estes mesões com estados 2S+1 LJ
do sistema quark-antiquark ud apropriado, especificando o valor do número
quântico principal n em cada caso.
R:

Nos é dado vários valores de J e P (J P = 0− , 1− , 0+ , ...). Como π +


é um mesão, sabemos que P = (−1)L+1 ⇒ daqui conseguimos L. De
J = L + S ⇔ S = J − L, podemos então deduzir S. Temos então o
necessário para encontrar 2S+1 LJ .
• 0− : P = −1 ⇒ L = 0, J = 0 = 0 + S ⇒ S = 0. Estado 1 S0
• 1− : P = −1 ⇒ L = 0, J = 1 = 0 + 4 ⇒ S = 1. Estado 3 S1
• 0+ : P = +1 ⇒ L = 1, J = 0 = 1 + S ⇒ S = −1. Estado 3 P0 .
• 1+ : P = +1 ⇒ L = 1, J = 1 = 1 + S ⇒ S = 0. Estado 1 P1 , e 3 P1
por algum motivo ????????
• 2+ : P = +1 ⇒ L = 1, J = 2 = 1 + S ⇒ S = 1. Estado 3 P2 .
5. Em geral, um mesão com spin J pode ter C = (−1)J ou C = (−1)J+1 e
P = (−1)J ou P = (−1)J+1 , havendo quatro possı́veis combinações de C
e P no total. Quais destas combinações podem ocorrer no modelo simples
de quarks? Faça a lista explı́cita dos valores JPC proibidos para J = 0, 1, 2
e 3.
R:
Mesão: qq ⇒ Spin S = 0 ∨ S = 1
Casos possı́veis:
• S = 0 ⇒ J = L:
– P = (−1)L+1 = (−1)J+1
– C = (−1)L+S = (−1)J
• S = 1, J = L ± 1:

35
– P = (−1)L+1 = (−1)J±1+1 = (−1)J
– C = (−1)L+S = (−1)J±1+1 = (−1)J
• S = 1, J = L:
– P = (−1)L+1 = (−1)J+1
– C = (−1)L+S = (−1)J+1

Ou seja, é proibido apenas o caso em que C = (−1)J+1 e P = (−1)J , i.e.


os casos J P C = 0+− , 1−+ , 2+− , 3−+ , ...
6. Os bariões com charme mais leves têm a composição de quarks cab com
momento angular orbital nulo L12 = L3 = 0, onde c é o quark c (charme)
e a e b podem ser qualquer um dos quarks leves u, d e s. Mostre que os
estados resultantes podem ser classificados em três famı́lias:
1+
• bariões 2 em que o par de quarks leves ab tem spin-0;
1+
• bariões 2 em que o par de quarks leves ab tem spin-1;
3+
• bariões 2 em que o par de quarks leves ab tem spin-1;
Faça a lista das combinações de quarks cab que ocorrem em cada uma
destas famı́lias e classifique-as em multipletos de isospin, deduzindo I3 a
partir das equações Y = B + S + C + B̃ + T e I3 = Q − Y /2.
R:
Se a = b, o spins de a e b têm de ser iguais, i.e. Sab = 1 (assim o spin
do trio poderia ser 1/2 ou 3/2). Por outro lado, se a ̸= b, o spin do par
pode ser 0 (spin do trio será 1/2) ou 1 (spin do trio pode ser 1/2 ou 3/2).
Temos então:
(a) a ̸= b, Sab = 0, S = 1/2: cud, cus, csd.
(b) a ̸= b, Sab = 1, S = 1/2: cud, cus, csd.
(c) a ̸= b, Sab = 1, S = 3/2: cud, cus, csd.
(d) a = b, Sab = 1, S = 1/2: cuu, cdd, css.
(e) a = b, Sab = 1, S = 3/2: cuu, cdd, css.
Agrupando como o problema pede, temos que
+
• item (a) = bariões 21 em que o par de quarks leves ab tem spin-0.
Examinando I3 , Q, B, S, C, B̃, T vemos que cus e csd fazem parte do
mesmo isodoublet, enquanto cud constitui um isosingleto.
+
• itens (b, d) = bariões 12 em que o par de quarks leves ab tem spin-1.
Examinando os números quânticos, vemos que aqui cud não é mais
um isosingleto, mas faz parte do isotripleto cud, cuu, cdd. Por outro
lado, cus, cds novamente formam um isodobleto e css um isosingleto.
Juntos, as 6 partı́culas neste estado formam um sexteto.

36
+
• itens (c, e) = bariões 32 em que o par de quarks leves ab tem spin-1.
Mesma coisa que o anterior, mas agora o sexteto é de spin 3/2.
7. Desenhe os diagramas de Feynman de ordem mais baixa para os seguintes
processos:
(a) a interação de um quark e um gluão para produzir um quark e um
fotão;
(b) a produção de um único bosão Z 0 na colisão de um protão com um
anti-protão;
(c) a aniquilação de um eletrão e um positrão para produzir um par de
bosões W.
R:
Ver figuras.

Figure 14: Problema 7(a)

Figure 15: Problema 7(b)

Figure 16: Problema 7(c)

37
8. Um colisionador pp com energias de feixe iguais é usado para produzir um
par de quarks top. Desenhe o diagrama de Feynman para este processo
que envolve apenas um gluão. Se os três quarks do protão (ou antiprotão)
trazem entre eles 50% da energia momento total do protão, calcule o mo-
mento mı́nimo do feixe necessário para produzir o par tt.
R:

9. (texto longo do caralho)


R:
Lembrando: estamos usando unidades naturais (nu), logo c = h̄ = 1.
Como W é a massa dos hadrões X, e E 2 = p2 + m2 ⇔ m2 = E 2 − p2 ,
temos
W 2 = EX 2
− p2X
Por conservação de energia e momento,

Ei = Ef ⇔ E + Ep → E ′ + EX ⇔ EX = (E − E ′ ) + Ep
pi = pf ⇔ p + Pp → p′ + pX ⇔ pX = (p − p′ ) + Pp
Então
W 2 = [(E − E ′ ) + Ep ]2 − [(p − p′ ) + Pp ]2
W 2 = (E − E ′ )2 + 2(E − E ′ )Ep + Ep2 − (p − p′ )2 − 2(p − p′ )Pp − Pp2
Mas em (nu) temos Q2 = (E − E ′ )2 − (p − p′ )2 e M 2 = Ep2 − Pp2 (de
E 2 = p2 + m2 ), então

W 2 = M 2 + Q2 + 2[(E − E ′ )Ep + (p − p′ )Pp ]

W 2 − Q2 − M 2 = 2[(E − E ′ )Ep + (p − p′ )Pp ]


Do enunciado, em (nu) temos 2M ν = W 2 + Q2 − M 2 . Não sei o que há
de errado no sinal do Q mas vamos fingir que é isso que temos no lado
esquerdo.
2M ν = 2[(E − E ′ )Ep + (p − p′ )Pp ]
Ep Pp
ν = (E − E ′ ) + (p − p′ )
M M
No referencial do protão, Pp = 0, ou seja Ep = Pp2 + M 2 = M 2 , então
2

Ep = M e
ν = (E − E ′ )
Agora, já que 2M ν = W 2 + Q2 − M 2 ,

Q2 Q2
X= = 2
2M ν W + Q2 − M 2
Então
0≤X≤1

38
6 Weak Interactions: Quarks and Leptons
• W + , W − and Z 0 are much more massive than gluons (and photons obvi-
ously), so the interactions are very short range.
• Masses:
MW = 80.40GeV /c2 , MZ = 91.19GeV /c2

• Such masses give the range:

RW ≈ RZ ≈ 2 × 10−3 f m

• At very high energies, the weak and EM interactions are the same. At low
energies, however, they can be clearly separated.

6.1 Charged Current Reactions


6.1.1 W ± -lepton interactions
• Basic vertices:

Figure 17: The two basic vertices for W ± -lepton interactions.

• Each vertex is characterised by the same dimensionless strength parameter


αW , independent of which lepton type is involved.
• Although the basic vertices cannot occur by themselves in nature because
of energy conservation, there are two exceptions, which do conserve energy
and momentum:

νl + l + → W + , W − → l− + ν l

• Remember universality: Weak and EM interactions are the same for


e, µ, τ , provided their mass differences are taken into account.

39
• We can find the rate of νe → e− + W + and W − + νe → e− using the
method of dimensions:

Γ(W → eν) ≈ 0.230 ± 0.008GeV

Γ(W → eν) ≈ αW MW ≈ 80αW GeV


αW = 0.0043 ± 0.0002

6.1.2 Lepton-quark symmetry


• The weak interactions of hadrons are understood in terms of basic pro-
cesses in which W ± are emitted/absorbed by the hadrons’ constituent
quarks. These can give rise to semileptonic processes or purely hadronic
processes, the latter being much harder to understand, for which reason
we shall ignore it for now. Also, we initially focus only on the u/d and c/s
quark generations, since they form the overwhelming majority of known
hadrons.

• The weak interactions of quarks are best understood in terms of two ideas:
lepton-quark symmetry and quark mixing.
• Lepton-quark symmetry: In its simplest form, it asserts that the two
generations of quarks u/d, c/s and the two generations of leptons νe /e− ,
νµ /µ− have identical weak interactions; i.e. one obtains the basic W ± -
quark vertices by making the replacements:
– νe → u
– e− → d
– νµ → c
– µ− → s
In the basic W ± -lepton vertices, leaving the coupling constant gW un-
changed.
This way, one obtains the vertices of the Figure 18, with coupling constants

gud = gcs = gW (ignoring quark mixing)

We can use this symmetry to investigate lepton number conservation in a


decay, for example. Then we can know whether it is possible. But we also
have to take quark mixing into account.
• Quark mixing: according to this idea, the d and s quarks participate in
the weak interactions via the linear combinations

d′ = d cos θC + s sin θC

s′ = −d sin θC + s cos θC

40
Figure 18: The W ± quark vertices obtained from lepton–quark symmetry when
quark mixing is ignored.

Figure 19: The ud′ W vertex and its interpretation in terms of udW and usW
vertices.

Where θC = Cabibbo angle. So in addition to the four vertices of


Figure 18 with the couplings

gud = gcs = gW cos θC (Cabibbo-allowed decay: no quark mixing)

one has the vertices of Figure 20 with the couplings

gus = −gcd = gW sin θC (Cabibbo-suppressed decay: with quark mixing)

• To find θC one must simply compare two reactions that involve them, for
example:
Γ(A → B) g2
∝ us 2
2 = tan θC
Γ(C → B) gud
gus /gud = tan θC = 0.232 ± 0.002
θC = 13.1 ± 0.1 deg

• Cabibbo Supressed decays: Decays that involve the couplings gus =


−gcd = gW sin θC , ”supressed” because their rates are typically reduced
by a factor of order
2 2 2 2
gus /gud = gcd /gcs = tan2 θC = 1/20

41
Figure 20: The additional vertices arising from lepton–quark symmetry when
quark mixing is taken into account.

6.1.3 W boson decays


• We know, from the fundamental Weak vertices, that W + can decay as
W + → l+ + νl . Lepton-quark symmetry with quark mixing imply that it
can also decay to quark-antiquark pairs
W ± → ud′ , cs′
We also have
ud′ = ud cos θC + us sin θC
so that ud and us will occur with relative probabilities sin2 θC and cos2 θC ,
respectively.
• In the rest-frame of the decaying W + , the quark and antiquark are emitted
back to back with high energies E ≈ MW /2 ≈ 40GeV . They fragment
into hadron jets, so that the observed decay reaction
W + → hadrons
is dominated by the two step mechanism of Figure 6.1.3.
• Relative rates for quark decays W ± → ud′ , cs′ and the leptonic decays
W + → l+ + νl are easily estimated in the good approximation that the
final state lepton and quark masses are neglected:
1 1
Γ(W + → ud′ ) = Γ(W + → cs′ ) = Γ(W + → e+ νe )
3 3
since the mechanisms of the reactions are identical, but the qq pairs can
be produced in 3 colour states with equal probabilities.
Furthermore, universality gives
Γ(W + → e+ νe ) = Γ(W + → µ+ νµ ) = Γ(W + → τ + ντ )
Since these are the only first-order weak decays possible, and since there
are two quark combinations (W ± → ud′ , cs′ ) contributing to the hadron
decays (W + → hadrons), we arrive at the branching ratio
B(W + → hadrons) ≈ 2/3
Where, for each of the three leptonic decay modes,
B(W + → l+ νl ) ≈ 1/9 (l = e, µ, τ )

42
Figure 21: The dominant mechanism for the decay of W ± bosons into hadrons.

6.2 The Third Generation (t/b quarks)


6.2.1 More quark mixing
• First let’s rewrite d and s mixing in the matrix form
 ′   
d cos θC sin θC d
=
s′ − sin θC cos θC s

• Now we include the third generation t/b:


 ′   
d Vud Vus Vub d
s′  =  Vcd Vcs Vcb  s
b′ Vtd Vts Vtb b

Where the CKM matrix Vαβ (α = u, c, t, β = d, s, b) must be unitary to


ensure that d′ , s′ and b′ are orthonormal single-quark states, like d, s, b.
• We then apply lepton-quark symmetry to the doublets u/d′ , c/s′ and t/b′
in order to obtain values for the αβW couplings gαβ . The argument is
analogous to the one used for just the two generations, and gives

gαβ = gW Vαβ (α = u, c, t; β = d, s, b)

• If we neglect mixing between b and (d, s), we get a useful approximation


   
Vud Vus Vub cos θC sin θC 0
 Vcd Vcs Vcb  ≈ − sin θC cos θC 0
Vtd Vts Vtb 0 0 1

43
which is equivalent to the first equation in the subsection together with

b′ = b

• The dominant decays of the b quark: Figure 6.2.1.

Figure 22: The dominant decays of the b quark to lighter quarks and leptons.

• Thus we have coupling constants

gub = gcb = gW

• Experimentally, b decays can be inferred from the decays of their hadrons,


and imply
τb ≈ 10−12 s
which is very long compared with what one would expect if the couplings
were not suppressed (i.e. if gub = gcb = gW )
Using the approximation of neglecting b and (d, s) mixing, the decay would
be  5
1 mτ
τb ≈ ≈ 10−15 s
N mb
Which also gives
ττ ≈ 3 × 10−13 s
Furthermore, a more thorough analysis yields

|Vub |2 ≈ 2 × 10−5 |Vcb |2 ≈ 2 × 10−3

In other words, while the neglect of mixing between b and (d, s) is not
exact, it is a very good first approximation.

6.2.2 Properties of the top quark


• The dominant decays of the t quark: Figure 6.2.2
• The properties of the top quark differ markedly from those of the other
quarks because it is so much heavier.

44
Figure 23: The dominant decays of the t quark to lighter quarks and leptons.
But t → d and t → s are insignificant

• The only significant decay mode is

t → b + W+

Because
gtd = gts = 0
gtb = gW

• The lifetime is
τt =≈ 4 × 10−25 s
i.e. it is highly unstable. The other five quarks u, d, c, s, b have lifetimes
of order 10−12 s or more, meaning they have time to form hadrons. In
contrast, when top quarks are created, they decay too rapidly to form
observable hadrons. Instead, they decay to b and W + , which in turn
decays predominantly to either lighter quarks by

W + → q1 + q 2 (q1 q 2 = ud, us, cd, cs)

or to leptons by
W + → l + + νl (l = e, µ, τ )

• Furthermore, the quarks released in these decays are not seen directly, but
”fragment” into jets of hadrons.

45
6.3 Selection Rules in Weak Decays
• These rules can be derived from examining the quark structures allowed
by conservation laws. Remember: hadronic decays only involve hadrons
- thus they only involve quarks, and semileptonic decays involve hadrons
(thus quarks) and leptons (no quarks).
• S = strangeness, Q = !QUARK! charge.
• ∆S = 0, ∆Q = ±1 or ∆S = ∆Q = ±1 for semileptonic decays
• ∆S = 0, ±1, ∆Q = 0 for hadronic decays

7 Weak Interactions: Electroweak Unification


7.1 Neutral Currents (Z 0 ) and the Unified Theory
• The existence of Z 0 was predicted by the unified theory.
• Its Feynman diagrams contribute to the branching ratios, cancelling out
divergences (infinities) in higher-order W ± -exchange diagrams (which would
previously just be ignored, since higher order = more complex = lower
chance of happening).

7.1.1 The basic vertices


• Lepton vertices: All known neutral current interactions can be ac-
counted for in terms of the basic Z 0 -lepton vertices (Figure 7.1.1):

Figure 24: Basic lepton vertices

• Quark vertices: The corresponding quark vertices can be obtained using


lepton-quark symmetry and quark mixing. In other words, if we again con-
fine ourselves to the first two generations, the quark vertices are obtained
from the lepton vertices by making the replacements νe → u, νµ →

46
c, e− → d′ , µ− → s′ .
The possible lepton vertices for Z 0 are

νe νe Z 0 , νµ νµ Z 0 , e− e− Z 0 , µ− µ− Z 0

Thus the corresponding quark vertices are

uuZ 0 , ccZ 0 , d′ d′ Z 0 , s′ s′ Z 0

The last two include d′ and s′ , so

d′ d′ Z 0 = (d cos θC + s sin θC )(d cos θC + s sin θC )Z 0 =

= ddZ 0 cos2 θC + ssZ 0 sin2 θC + (dsZ 0 + sdZ 0 ) sin θC cos θC


and

s′ s′ Z 0 = (−d sin θC + s cos θC )(−d sin θC + s cos θC )Z 0 =

= ddZ 0 sin2 θC + ssZ 0 cos2 θC + (dsZ 0 + sdZ 0 ) sin θC cos θC


However, when we combine these two contributions, we get

d′ d′ Z 0 + s′ s′ Z 0 = ddZ 0 + ssZ 0

such that the four basic vertices may be replaced by the equivalent vertices

uuZ 0 , ccZ 0 , ddZ 0 , ssZ 0

The four vertices of Figure 7.1.1 conserve s and c, whereas the ”flavour-
changing” vertices ucZ 0 and dsZ 0 do not occur. Thus, we arrive at the
conclusion that...
• neutral current interactions conserve strangeness and charm, in
contrast to the charged current interactions.

7.1.2 The unification condition and the W ± and Z 0 masses


• For unification, two things must be satisfied: the unification condition and
the anomaly condition.
• Unification condition:
e
= gW sin θW = gZ cos θW
2(2ϵ0 )1/2
Where θW = weak mixing angle,

θW = MW /MZ (0 < θW < π/2)

and gZ = coupling constant that characterises the strength of the neutral


current vertices of Figures 7.1.1 and 7.1.1.

47
Figure 25: Basic quark vertices. a = u, d, s, ...

• Anomaly condition:
X X
Ql + 3 Qa = 0
l a

Where Ql = electric charge of the leptons, Qa = electric charge of the


quarks.
On substituting for the six known leptons l and six known quarks a, one
finds the condition is satisfied.
• The unification condition was historically used to predict the masses of
the bosons W ± and Z 0 :
√ 2
2 2gW πα
– MW = GF = √
2GF sin2 θW
= 78.3 ± 2.4GeV /c2
πα
– MZ2 = √
2GF sin2 θW cos2 θW
= 89.0 ± 2.0GeV /c2

• Weak mixing angle:

sin2 θW = 0.2315 ± 0.0001

7.1.3 Electroweak reactions


• In any process in which a γ is exchanged, a Z 0 can be exchanged as well.

48
• At low energies, i.e. at energy and momentum transfers that are small
compared to the Z 0 mass, the Z 0 -exchange contribution can be neglected,
and the reaction is purely electromagnetic to a high degree of accuracy.
• At high energies, Z 0 -exchange contributions become comparable with pho-
ton exchange, thus we are dealing with electroweak processes.
• Vertices:

Figure 26: Summary of the Z 0 and γ couplings to leptons and quarks in the
unified electroweak theory, where l = e, µ, τ and a = u, d, s, ...

• Example: this is illustrated by the reaction e+ + e− → µ+ + µ−

– Cross-section of the reaction: We can see that at low-energies, the γ-


exchange dominates and the energy falls of as E −2 . At about 25GeV ,
Z 0 begins to make a significant contribution. At still higher energies,
we have a peak at center-of-mass energy

ECM = 2E = MZ

corresponding to the Z 0 mass. This corresponds to the formation of


physical Z 0 bosons in the process

e+ + e− → Z 0

followed by the subsequent decay

Z 0 → µ+ + µ−

The width of the peak gives the total decay rate,

ΓZ = 1/τZ = 2.49GeV

49
+
Figure 27: Total cross-section for the reaction e+ + eßµ +µ as a function of the
total centre-of-mass energy ECM . The dashed line shows the extrapolation of
the low-energy behaviour σγ in the region of the Z 0 peak.

– γ-exchange contribution:
α2
σγ ≈
E2
– Z 0 -exchange contribution:
σZ ≈ G2Z E 2

– The ratio of the two gives


σZ G2 E 4 E4
≈ Z2 ≈ 4
σγ α MZ

7.1.4 Z 0 formation: how many neutrinos are there?


• The Z 0 produced in e+ + e− → Z 0 will not only decay to muon pairs, but
also to other final states allowed by the appropriate conservation laws.
• The decays seen in experiment are
e+ + e− → l + + l − (l = e, µ, τ )
and
e+ + e− → hadrons
where a sum over all possible hadrons is implied.

50
• Bret-Wigner formula: describes the cross-section due to the formation
and decay of Z 0 bosons.

12πMZ2 Γ(Z 0 → e+ + e− )Γ(Z 0 → X)


 
σ(e+ + e− → X) = 2 2
ECM (ECM − MZ2 )2 + MZ2 Γ2Z

• Using experimental data, one finds

MZ = 91.188 ± 0.002GeV /c2

ΓZ = 2.495 ± 0.002GeV
These two decay rates are just over 80% of the total:

Γ(Z 0 → l+ + l− ) = 0.0840 ± 0.0009

Γ(Z 0 → hadrons) = 1.74 ± 0.002GeV


The rest are states with just neutrinos produced.
• In unified theory, the only Z 0 that can arise from the basic vertices are

Z 0 → νl + ν l

Z0 → l + l
Z0 → q + q
In the last one, the quark pair isn’t seen directly, but instead fragments
into two or more jets of hadrons that are observed in the final state.
• On adding up the various decays we thus obtain

ΓZ = Γ(Z 0 → hadrons) + 3Γ(Z 0 → l+ + l− ) + Nν Γ(Z 0 → νl + ν l )

The total neutrino contribution is therefore

Nν Γ(Z 0 → νl ν l ) = 0.499 ± 0.004GeV

The decay rate to neutrino pairs cannot, of course, be measured directly,


but can be calculated from its Feynman diagram

Γ(Z 0 → νl ν l ) = 0.166GeV

• (...) We thus see that, within the SM, of which the unified electroweak
theory is a part, there can only be three generations of leptons and quarks
if neutrinos are assumed to be light compared with the Z 0 mass.

51
7.2 Gauge Invariance and the Higgs Boson
• Gauge invariance is a fundamental symmetry associated with theories in
which the force carriers are spin-1 bosons.
• There are different forms of gauge invariance, corresponding to the differ-
ent interactions of particle physics.
• Their common feature is that the parameters of the corresponding sym-
metry transformations are allowed to vary with position (⃗r, t) in space and
time.
• Principle of minimal gauge invariance, or gauge principle: approach in
which the form of the interaction is inferred by adding the minimal inter-
action terms needed to make an equation of motion gauge-invariant.

7.2.1 Unification and the gauge principle


• In order to agree with observations, we must unite EM and weak forces.
To do this we regard the photon γ and the neutral boson Z 0 as mixtures
of the W 0 with another neutral boson B 0 , i.e.
γ = B 0 cos θW + W 0 sin θW
Z 0 = −B 0 sin θW + W 0 cos θW
• Those transformations lead to the vertices
e− → e− B 0 , νe → νe B 0
analogous to e− → e− γ.
• In this way, one arrives at a unified theory of electroweak interactions that
is gauge-invariant. The agreement between the predictions and experiment
are great, bla bla bla. However, there is one remaining problem, to which
we now turn.

7.2.2 Particle masses and the Higgs field


A PARTIR DAQUI JÁ NÃO ENTENDO REALMENTE NADA. TERMINAR
CAP. 1, E LER CAP. 2.

8 Discrete Symmetries: C, P , CP and CP T


8.1 P violation, C violation and CP conservation
8.1.1 P violation
• P violation was first proved to exist in Weak interactions by analysing the
reaction
60
Co → 60 N i + e− + ν e

52
There was a ”forward-backward decay assymetry”, i.e. the fact that fewer
electrons are emitted in the forwards direction than backwards with re-
spect to the spins of the decaying nuclei. This violates parity because
the parity operator switches the emitted electrons’ momenta as p → −p
but keeps the nuclear spin the same, suggesting that we would see 50/50
electrons being emitted in each direction in relation to the nuclear spin
hemisphere. Yet that is contradicted by experiment.

60
Figure 28: Effect of a parity transformation on Co decay.

8.1.2 C violation (and another P violation)


• Discovered with muon decay symmetries. Consider the muon decays

µ− → e− + ν e + νµ µ+ → e+ + νe + ν µ

• In the rest frame of the decaying particle these were found to be of the
form  
1 ξ±
Γµ± (cos θ) = Γ± 1 − cos θ (1)
2 3
Where
– θ = angle between muon spin direction and direction of outgoing
electron/positron
– ξ± = assymetry parameters
−1
R +1
– Γ± = total decay rates, or inverse lifetime τ± = −1
d cos θΓµ± (cos θ) =
Γ±
• C transforms particles → antiparticles, so they transform the µ− decay
into the µ+ decay and vice-versa. Thus C-invariance implies that the rates
and angular distributions for these decays should be the same, i.e.

Γ+ = Γ− ξ+ = ξ− (C-invariance)

53
Figure 29: Effect of a parity transformation on the muon decays. The short
thick arrows indicate the direction of the muon spin, while the long arrows
show the direction of the electron’s momentum.

• P preserves the identities of the particles, but reverses their momenta,


while keeping the spins unchanged. Hence P -invariance implies
Γµ± (cos θ) = Γµ± (− cos θ) (P-invariance)
and substituting in 1 we get
ξ± = 0 (P-invariance)

• Experimentally, we know that


ξ− = −ξ+ = 1.00 ± 0.04
which shows that both C and P invariances are violated.

8.1.3 CP conservation
• However, experimentally we also know that Γ+ = Γ− , i.e. µ+ and µ−
do in fact have equal lifetimes. Why do they have the same lifetime if
C-invariance is violated? Because of CP conservation.
• CP -conservation: the weak interaction is invariant under a CP trans-
formation, even though both C and P are separately violated. If we apply
the CP operator to muon decays, the P changes θ → π −θ as before, while
the C operator changes the particle to antiparticle. Hence, CP invariance
alone implies
Γµ+ (cos θ) = Γµ− (− cos θ)
Substituting into 1 we get
Γ+ = Γ− ξ+ = −ξ− (CP-invariance)

• The weak interactions of quarks do not respect CP invariance, although


it is often a very good approximation, and the observed violations are so
far confined to the decays of neutral kaons and B mesons.

54
8.1.4 Left-handed neutrinos and right-handed antineutrinos
• Helicity States: the spin is quantised along the direction of motion of
the particle, rather than some arbitrarily chosen ”z-direction”.

Figure 30: Helicity states of a spin- 12 particle: long thin arrows = momenta of
the particles, shorter thick arrows = their spin.

– Right-handed or positive helicity: spin along direction of motion + 21


– Left-handed or negative helicity: spin along direction of motion − 12

.
• Only left-handed neutrinos νL and right-handed antineutrinos ν R are ob-
served in nature.
– This violates C-invariance, which requires neutrinos ν and antineu-
trinos ν to have identical weak interactions, since the C operator
converts a particle to its corresponding antiparticle.
– This also violates P -invariance, which requires the states νL and νR
to have identical weak interactions, since the P operator reverses
the momentum while leaving the spin unchanged, thus converting a
left-handed neutrino to a right-handed neutrino.
– This does not violate CP -invariance, since the CP operator converts
a left-handed neutrino do a right-handed antineutrino.
• The fact that only left handed-neutrinos and right-handed antineutrinos
are observed in nature extends to all fermions in the ultra-relativistic limit
(with v → c). For neutrinos specifically, that is always a good approxima-
tion. For other particles, it is only a good approximation for large E.
• You can use momentum conservation, together with the fact above, to
predict the helicity states of certain reactions.

55
8.2 ETC
• Amplitude Equation:

−g 2 h̄2
F (⃗q2 ) = 2 c2
|⃗q|2 + MX

Where
– MX = mass of exchanged particle
– g = appropriate coupling constant
– The amplitudes of the EM and Weak interactions only become of
comparable size when |⃗q|2 ≈ MX
2 2
c , i.e. at high energies.

8.3 SÉRIE 4
1. Desenhe o diagrama de Feynman de ordem mais baixa a nı́vel de quarks
para o declı́nio Λ → p + π − . Se a natureza decidisse passar par o dobro
a constante de interação fraca e diminuisse a massa do bosão W por um
fator de 4, qual seria o efeito na taxa deste declı́nio?
R:

2
−gW h̄2 g2
amplitude = 2 ∝ W2
|q|2 c2− MW c 4 MW
onde se usou o fato que q 2 c2 << MW
2 4
c . Temos então
4
gW
Γ ∝ amplitude2 ∝ 4
MW

(2gW )4 4
gW
= 4096
( 14 MW )4 MW4

56
2. Use a universalidade da interação fraca e a simetria quarks-leptões para
estimar as razões de bifurcação (taxa/taxa total) para (a) os declı́nios
b → c + e− + ν e (onde os quarks b e c estão em estados ligados hadrónicos)
e para (b) os declı́nios τ − → e− + ν e + ντ . Ignore os estados finais que
sejam “Cabibbo-suppressed” em relação aos modos leptónicos.
R:
Para saber as razões de bifurcação, precisamos considerar todos os decai-
mentos possı́veis para os reagentes em questão. Assim, podemos saber
a probabilidade do decaimento especı́fico dado em comparação com os
outros possı́veis.
(a) Para o decaimento,
b → c + e− + ν e
Também temos os decaimentos

b → c + µ− + ν µ

b → c + τ − + ντ
Por universalidade leptónica, os três têm a mesma taxa de decaimento
Γ.
Além disso, podemos ter decaimento da forma

b→c+X

em que X é um hadrão permitido pelas leis de conservação (então


tem de ter carga −1).
Os únicos vértices permitidos sem quark-mixing (ou seja, ignorando
estados Cabbibo-supressed) são u com d e c com s, isto é, udW , udW ,
udW , udW e, similarmente, csW , csW , csW e csW . Mas destes, os
únicos que têm carga −1 são udW e csW . Cada um destes vértices
tem peso estatı́stico 3, pois os quarks existem em 3 estados de cores.
Temos então os 3 estados dos decaimentos leptónicos, e 2 × 3 = 6
estados dos decaimentos hadrónicos. Ao todo, 9 estados possı́veis.
Então,
1
B(b → c + e− + ν e ) =
9
(b) Para o decaimento
τ − → e− + ν e + ντ
Também temos o decaimento

τ − → µ− + ν µ + ντ

Obviamente não consideramos τ − → τ − + ν τ + ντ pois este decai-


mento não muda nada.
Além disso, temos os decaimentos hadrónicos

τ − → ντ + X

57
em que X é um hadrão permitido pelas leis de conservação, com
carga −1.
A princı́pio, ignorando os vértices Cabibbo-supressed, os únicos vértices
permitidos com carga −1 são X = ud∨cs. Mas, como ms +mc > mτ ,
o cs não é permitido, pelo que temos apenas o decaimento

τ − → ντ + u + d

Que contribui ×3, devido às 3 cores possı́veis.


Temos então 2 decaimentos leptónicos e 3 decaimentos hadrónicos
possı́veis, para um total de 5. Então,
1
B(τ − → e− + ν e + ντ ) =
5
3. A partir do exercı́o 2, estime a vida média dos declı́nios semileptónicos do
quark c: c → s + e+ + νe . Obter as vidas-médias dos seguintes declı́nios:

• D + → e+ + X
• D 0 → e+ + X
• Λ+ +
C →e +X

Compare com a estimativa anterior. R:

4. Considere os seguintes processos e diga se são permitidos na interação


eletromagnética, e quais são permitidos na interação fraca através da troca
de um único bosão W + , W − ou Z 0 .
R:

(a) K + → π 0 + e+ + νe
K + = us, π 0 = uu ∨ dd. Já não pode ser EM pois muda a estranheza
S. Se π 0 = uu, podemos fazer um diagrama com o vértice suW + ,
que é permitido com quark-mixing. Também podemos usar as re-
grinhas de ∆S e ∆Qquark (dá no mesmo pois as regras são tiradas
dos vértices): é um decaimento semileptónico com ∆S = ∆Q = −1,
logo permitido.
(b) K + → π + + e+ + e−
K + = us, π + = ud. Não pode ser EM pois muda estranheza S. Nesse
caso terı́amos de ter um vértice sdW , que não é possı́vel. Também
podemos descobrir isso com as regrinhas: ∆S = −1, ∆Q = 0, logo
não é permitido.
(c) Σ0 → Λ0 + e+ + e−
Σ0 = uds, e Λ0 = uds também. Então não há mudança na estran-
heza: pode ser EM (γ) ou Fraca Neutra Z 0 .

58
(d) Σ0 → n + νe + ν e
Σ0 = uds, e Λ0 = udd. Então tem-se s → d, o que já não é um
vértice permitido (sdW ), nem com quark mixing. Usando as regrin-
has: ∆S = +1, ∆Qquark = 0, proibido.
5. Negligenciando a massa do eletrão, o espetro de energia para os eletrões
emitidos no declı́nio do muão é dado por (expressão gigante). Qual é a
energia mais provável para o eletrão? Desenhe um diagrama mostrando os
momentos das 3 partı́culas do estado final e as suas helicidades para o caso
em que Ee ≈ mµ c2 /2 . Mostre também o spin do muão. Integre o espetro
de energia para obter um expressão para a largura total do decaimento do
muão. A partir daı́ calcule a vida média do muão em segundos. Dados:
GF /(h̄c)3 = 1.166 × 10−5 GeV −2
R:
A interação em questão é

µ− → e− + νe + ν µ

Estado de energia mais provável é no topo da curva, i.e. quando a derivada


da expressão do espectro é zero (o espectro dá as probabilidades):
 
d dω 1
= 0 ⇔ Ee = mµ c2
dEe dEe 2

Quando Ee ≈ 21 mµ c2 , o eletrão tem energia máxima ⇒ os dois neutrinos


devem voltar juntos. Além disso, a interação fraca só produz partı́culas
canhotas (spin e mom. linear opostos) e antipartı́culas destras (spin e
mom. angular na mesma direção). Temos então Integrando o espectro de

energia, obtemos o Γ da reação (área embaixo da curva da energia):


mµ c2
G2F (mµ c2 )5
Z 2
Γ = (...) (...)dEe = ≈ 3 × 10−19 GeV
0 192π 3 (h̄c)6

O tempo de vida é, então,



τ= ≈ 2.2 × 10−6 s
Γ

59
6. Analise os declı́nios do pião π + → µ+ + νµ e π + → e+ + νe e sob o ponto
de vista da helicidade das partı́culas do estado final. Qual dos declı́nios é
mais provável? Justifique.
R:
NAO SEI SE ESTÁ CERTO!
Do Particle p. 286, falando exatamente sobre estes decaimentos: ”For
the case of a positive muon this is unimportant, since it is easy to check
that it recoils nonrelativistically and so both helicity states are allowed.
However, if a positron is emitted it does recoil relativistically.”
Sabemos, de certeza, que
• Considerando o referencial de inércia do pião, o l+ e νl terão de
ser emitidos em direções opostas devido à conservação de momento
linear.
• Os neutrinos sempre devem ser destros.
• Os positrões e+ tem v → c, pela citação acima, e sabemos que no
limite relativista os positrões devem ser sempre destros.
• O mesão π + pode ter spin nulo ou 1, enquanto os léptons tem spin
±1/2. Então para preservar o spin, podemos ter os spins dos léptons
se cancelando (= 0) ou somando (= 1).
A citação acima também diz que os muões positivos não precisam neces-
sariamente ser destros, pois não são relatistas na emissão. Sendo assim,
podemos deixa-los canhotos ou não, pelo que o decaimento π + → µ+ + νµ
pode ocorrer de duas formas. Por outro lado, o positrão tem de ser de-
stro, então esse decaimento só pode ocorrer de uma forma. Portanto, o
primeiro decaimento é mais provável.

7. Considere o declı́nio do Cobalto-60, observado em 1957 pela Madame Wu


com núcleos polarizados de Cobalto. Observa-se que o espetro angular

60
dos eletrões emitidos é consistente com a expressão:
v
I(v, θ) = 1 + α cos θ
c
em que v = módulo da velocidade dos eletrões e θ = ângulo entre a sua
direção e a direção do spin do Cobalto-60. Deduza qual o valor de α,
considerando eventos em que o eletrão é emitido na direção do spins dos
núcleos que decaem. Os spins dos núcleos Co60 e Ni60 são J = 5 e J = 4,
respetivamente. e o momento angular orbital pode ser desprezado.
R:
60
Co → 60 N i(J = 4) + e− + ν e
Para o spin J = 5 do Co ser preservado, é necessário que e− e ν e tenham
spin S = +1/2 cada (o momento angular orbital L é desprezado).
Como estamos considerando eventos em que o eletrão é emitido na direção
do spin do núcleo (θ = 0), e como já vimos que o spin do núcleo e do elétron
tem o mesmo sinal e portanto mesma direção, então logicamente o eletrão
tem de ser emitido na direção do spin do núcleo. Noutras palavras, o e−
há de ser destro, i.e. e−
R.
No entanto, como vimos na teoria, ”The fact that only left handed-
neutrinos and right-handed antineutrinos are observed in nature extends
to all fermions in the ultra-relativistic limit (with v → c)”, ou seja, quando
ve → c o eletrão tem de ser canhoto. Assim, nestas condições temos de
ver intensidade nula dos eletrões emitidos:

I(v = c, θ = 0) = 0

Ou seja,
1 + α = 0 ⇔ α = −1

9 Nuclear Properties
Bibliografia: cap. 3 e 5 do krane; cap. 4 do w.n. cottingham and d.a. green-
wood; slides.

9.1 Dispersão de Rutherford e dimensão do núcleo


• Dispersão de Rutherford: partı́culas α (núcleos de He) sendo atirados com
diferentes energias no quadrado de Pb. Ver figura 31
• O potencial entre o núcleo de He e o núcleo de Pb é

q 1 e2 ZHe ZP b
V =k = = 26.2M eV
r 4πϵ0 RHe+P b

Dessa forma se soubermos V podemos obter a dimensão do núcleo (RHe+P b )

61
Figure 31: .

• Para o mesmo ângulo (60 deg), varia-se a energia da α e se mede a inten-


sidade de α’s retrodispersos para cada energia. Quanto maior a energia,
mais podem ”vencer” a repulsão coulombiana que faz eles serem retrodis-
persos, logo a intensidade de retrodispersos para este ângulo diminui com
a energia. Eventualmente há uma queda rápida, quando atingem a energia
de repulsão V = 26.2M eV (é assim que sabemos o valor de V ). Ver figura
32

9.2 Difração de Eletrões e distribuição de carga nuclear


• Para detectar um objeto, o c.d.o. deve ser menor que as dimensões do
objeto; caso contrário a difração nos impede de vê-lo. Para os núcleos,
com diâmetro de 10fm, requer-se que λ < 10fm. p = 2πh̄ λ = 100MeV/c.
Já que essas partı́culas também são ondas, a difração delas faz com que o
primeiro mı́nimo apareça em θ = arcsin 1.22λ/D, tal que D é o diâmetro
do cı́rculo (buraco) através do qual a luz é difratada. Ver figuras 33 e 34.
• A distribuição de carga nuclear é ρ(x) = Zef (x) tal que f (x)d3 x = 1.
R

• dΩ
dσ dσ
 
exp
= dΩ teo
F (q 2 )
iqx
• Fator de forma F (q 2 ) = f (x)d3 x
R
e h̄

2
• Raio quadrático médio ⟨r2 ⟩ = −6h̄2 dFdq(q2 ) |q2 =0

• O resultado destes experimentos é claro - a densidade de carga central


nuclear ρ(x) é basicamente a mesma por todo o núcleo. Nucleões não se

62
Figure 32: .

congregam no centro do núcleo, mas têm uma distribuição bem constante


até a superfı́cie (do núcleo). A conclusão é a mesma para as massas.
Figura 35 mostra distribuição de carga para diferentes núcleos.
• Distribuição de carga nuclear: vemos os dados experimentais e descreve-
mos a curva que se encaixa melhor usando poucos parâmetros, nomeada-
mente o raio r (quando presumimos simetria esférica do núcleo, o raio é
um fator muito importante), a distância para qual a densidade se mantém
”constante” R, e distância a, que vai de R até o ponto em que a densidade
vira zero (ver figura 35). A expressão que melhor encaixa nos dados acaba
sendo
ρch (r) = ρ0ch /[1 + exp((r − R)/a]
O parâmetro a é basicamente igual para todos núcleos, mas o R aumenta
com A = Z + N .

• Relação entre densidade de carga ρ e número de massa A: ρ ∝ A


R3 . Ver
figura 36

9.3 Massas e energies de ligação de núcleos em seu estado


fundamental
• De Einstein E = mc2 , a energia de ligação que liga todos os protões e
neutrões do núcleo B(Z, N ) é

B(Z, N ) = [Zmp + N mn − mnuc (Z, N )]c2

em que mnuc é a massa de repouso do núcleo.

63
Figure 33: .

• Experimentalmente, são as massas de átomos com eletrões, ao invés das


massas de núcleos em si, que são medidos diretamente. Por E = mc2
temos
B(Z, N ) + be = [Z(mp + me ) + N mn ]c2
onde be é a energia de ligação dos elétrons. A contribuição dos elétrons é
negligenciável.

9.4 A fórmula de massa semi-empı́rica


• Nós já vimos que a densidade de massa nuclear é aprox. constante, e que
os núcleos tem uma superfı́cie bem definida (a pequeno). Sob esse ponto
de vista podemos ver o núcleo como uma ”gota de lı́quido”.
• A fórmula semi-empı́rica de massa expressa isso matematicamente, e ac-
erta com muita precisão exceto para os núcleos mais leves. Há muitas
formas, mas a nossa será

(N − Z)2 δ
B(Z, N ) = av A − as A2/3 − aC Z(Z − 1)A−1/3 − asym − 1/2
A A

64
Figure 34: .

Figure 35: .

• Termo de volume av A: representa uma energia proporcional ao volume no


sentido de ser a energia de cada nucleão com seus vizinhos, como a energia
coesiva de um lı́quido simples.
• Termo de Superfı́cie: −as A2/3 termo de correção para os nucleões da
superfı́cie, pois só tem vizinhos de um lado.
• Termo de Coulomb: −aC Z(Z −1)A−1/3 correção que considera a repulsão
entre os protões.
2
• Termo de assimetria: −asym (N −Z)A correção para a assimetria se o número
de neutrões for diferente do número de protões (olhe o N − Z)

65
Figure 36: .

• Termo de emparelhamento: − A1/2


δ
empı́rico

+11.2MeV
 N, Z ı́mpar
δ = 0MeV A ı́mpar

−11.2MeV N, Z par

av = 15.835MeV
as = 18.33MeV
ac = 0.714MeV
asym = 23.20MeV

• Contribuição de cada termo: ver figuras 37 e 38. Os dados experimentais


são a linha com x.

Figure 37: .

66
Figure 38: .

Figure 39: .

• Da figura 39 vemos que a relação da energia de ligação por nucleão B/A


com o número de massa A = Z + N forma uma parábola. A figura
40 mostra como chegamos ao mı́nimo desta parábola: da expressão para
ma (Z, N ), junto com A = Z + N , obtemos uma expressão ma (Z, A).
Simplificamos esta expressão definindo alguns parâmetros α, β e γ para
”resumir” a equação. Vemos que sai a expressão de uma parábola y =
a + bx + cx2 :
ma (Z, A)c2 = α − βZ + γZ 2
Para achar o mı́nimo basta fazer
∂ma
=0
∂Z
e encontramos a expressão para o Zmin
β
Zmin =

• Note-se pela figura 39 que para núcleos com A ı́mpar, só há 1 parábola.
Mas para A par, há 2 parábolas, uma para Z e N ambos ı́mpares, outra

67
Figure 40: .

para Z e N ambos pares. Só existem 4 núcleos estáveis com N e Z ambos


ı́mpar. Existem 167 núcleos estáveis para

9.5 Nuclide Charts


• Nuclide: a type of atom specified by its Z, A and its energy state. Rep-
resented as A
Z XN .

• Only ∼ 10% of known nuclides are stable: usually the ones with Z ∼ N
• Nuclide chart: plot number of protons Z vs. number of neutrons N in the
nucleus. Since the quantities are discrete, the nuclides will be little boxes.

Figure 41: .

• Each nuclide box contains basic properties of the nuclide.


• The transitions due to proton decay, neutron emission, α decay etc. are
available in figure 43
• The origin of the elements is in figure 44

9.6 Indicadores da existência de estrutura em núcleos


Veremos cada um dos indicadores em detalhe nas subseções seguintes.

68
Figure 42: .

Figure 43: .

9.6.1 Distribuição de abundâncias no sistema solar


• Ver figura 45

9.6.2 Energias de separação S2p e S2n


• The shell model was remarkable for explaining the relationship between
electrons and their nuclei. In the shell model, we fill the shells with elec-
trons in order of increasing energy, being consistent with the Pauli prin-
ciple. We obtain an inert core of filled shells and some number of valence
electrons. The model then assumes that atomic properties are determined
primarily by the valence electrons. In particular, we see regular/smooth
variations of properties within a subshell, but sudden/dramatic changes
when we fill one subshell and go to the next.

• Because of the success of the shell model for electrons, we try to apply
the same model for the nucleons. Figure 46 shows the separation energy
(equivalent to the ionisation energy for electrons) required to separate the
nucleons, as a function of the number of nucleons. S2p is the energy to sep-
arate 2 protons and S2n is the energy to separate 2 neutrons. We see that
the sudden changes occur at the same ”magic number” of nucleons, as for
the ”magic number” of electrons for electron shells: 2, 8, 20, 28, 50, 82, 126.
This is evidence that nucleons also follow a shell structure, like electrons.

69
Figure 44: .

Figure 45: .

• But how can the nucleons move like electrons? Two problems arise:

– In the case of electrons, their energies are determined by the electric


potential from the nucleus, an external agent. In the nucleus, there
is no such external agent; the nucleons move in a potential that they
themselves create.
– The electrons can move in their spatial orbits being relatively free
of collisions with other electrons. Nucleons, however, have a large
diameter compared to the size of the nucleus, so how can they move
in well defined orbits when a single nucleon can make many collisions
during each orbit?
• The solution is in the Pauli principle. When two nucleons collide, they
transfer energy with each other. But if all the energy levels are filled up
to the level of the valence nucleons, the only place a nucleon can go after
receiving energy, is to the valence shell. But if the energy gained from the
collisions is not nearly high enough for this (which is the case) then the

70
Figure 46: .

nucleons cannot collide - they can indeed orbit as if they were transparent
to one another! No classical analog, you have to accept

9.6.3 Energia de Partı́culas α emitidas por isótopos de Sn


• A figura 47 mostra as energias de partı́culas α (ou seja 42 He) emitidas
por um núcleo de Sn. Dentro de um shell, vê-se que as energias vão
diminuindo, até passar para o próximo shell onde a partı́cula α é emitida
com mais energia, haja vista que está no inı́cio da nova camada de valência
e logo pode ser arrancada mais facilmente.

Figure 47: .

71
9.6.4 Energia do estado 2+ para núcleos com Z e N pares
• Tal como na teoria dos gases, podemos examinar os núcleos mais pesados
em termos de suas propriedades gerais, ao invés de tentar construir uma
teoria a partir dos fundamentos. Ao fazer isso, vemos que há propriedades
comuns a todos os núcleos, as propriedades coletivas.
• Uma dessas propriedades é a energia do primeiro estado excitado de 2+ ,
que é anómala no sentido de ser ”mais baixa do que deveria ser”. A
Figura 48 mostra a energia desse estado, E(2+ ), versus o número de massa
A = Z + N . Vê-se novamente a trend dos números mágicos.

Figure 48: .

O resultado de todos esses indicadores é: estabilidade especial para núcleos


com número de protões/neutrões igual a 2, 8, 20, 28, 50, 82, 126, os ”números
mágicos” que também se encontram na teoria de shells dos eletrões.

9.7 Potencial no Modelo de Camadas


• The first step in developing our nucleon shell model is the choice of po-
tential, which determines the energies of each shell.
• The infinite well is not a good approximation to the nuclear potential
because i) it requires an infinite amount of energy to separate a neutron
or proton from the nucleus; ii) the nuclear potential doesn’t have a sharp
edge, but falls off smoothly to zero beyond the mean radius R.
• The harmonic oscillator is not a good approximation to the nuclear po-
tential either, because in this case the potential doesn’t fall off sharply
enough.

• Naturally we pick a middle ground between the two (the Woods-Saxon


parametrisation):
V0
V (r) = −
1 + exp [(r − R)/a]
Where R = 1.25A1/3 and a = 0.524f m, as seen before, and the well-depth
is adjusted to match the experimental separation energies: V0 ≈ 50M eV .

72
• With Woods-Saxon, filling the shells in order we get the first magic num-
bers 2,8, and 20, but the higher ones are different, in disagreement with
the truths seen before. Thus, we need to adjust this intermediate model
until it matches with the experimentally observed magic numbers.
• Success was achieved in 1949 by Mayer, Jensen et al. by including a
spin-orbit potential, which gives the proper separation of subshells.
• In atomic physics (which is about nucleus + electrons, unlike nuclear
physics which is just the nucleus), the spin-orbit interaction (which causes
the fine structure of spectral lines) comes about because of the EM interac-
tion of the electron’s magnetic moment with the magnetic field generated
by its motion about the nucleus. Such EM interactions are too weak to
produce the energy separations required so they make no difference.
• Thus we adopt the concept of a nuclear spin-orbit interaction of the same
form, but certainly not EM in nature.
• Nuclear spin-orbit force: interaction, not EM in nature, of the nucleon’s
magnetic moment with the magnetic field generated by its motion around
the nucleus it is in. Has the form

Vso (r)l · s

The form of Vso is not too important, its the l · s that gives the proper
energy level separation.
• We label the states according to their total angular momentum

j =l+s

just like in atomic physics.

9.8 Configuração do estado fundamental dos núcleos


• Z par, N par (A par) todos os nucleões acoplados em pares ⇒ J = 0, π = 1

• A ı́mpar ⇒ sobra um nucleão desemparelhado ⇒ J = j do nucleão de-


semparelhado, π = π do nucleão desemparelhado
• Z ı́mpar, N ı́mpar ⇒ não existem predições para J, e π = produto das π
dos nucleões desemparelhados.

• Paridade:
π = ΠA
i=1 (−1)
li
πp = πn = 1

73
10 Decaimentos Alfa, Beta, Gama
10.1 Decaimento Alfa
• Processo no qual o núcleo desintegra-se emitind uma partı́cula α = 42 He:
A
Zx → A−4 4
Z−2 Y + 2 He

• Do termo de repulsão coulombiana da fórmula semi-empı́rica, −aC Z(Z −


1)A−1/3 , vemos que este termo ∝ Z 2 .
• A partı́cula α é um estado muito estável

• De E = mc2 , o Q da reação (valor da energia absorvida/emitida na reação)



Q = [M (A, Z) − (M (A − 4, Z − 2) + M (4, 2))]c2

• O decaimento só é possı́vel se Q > 0, senão p ≤ 0 dos produtos. Ver figura


49.

Figure 49: .

• Na figura 50, os núcleos com...


– Q < 0 MeV: absolutamente estáveis contra decaimento alfa.
– 0 < Q < 4 MeV: tempo de vida de bilhões de anos, em prática
estável.
– Q > 4 MeV: núcleo instável, decai α.

Ver figura 50.


• Estimativa do valor Q utilizando a fórmula semi-empı́rica:

Q = [M (A, Z) − (M (A − 4, Z − 2) + M (4, 2))]c2

Q = B(A − 4, Z − 2) + B(4, 2) − B(A, Z) = Bα − ∆B


Em que

74
Figure 50: .

– ∆B = B(A, Z) − B(A − 4, Z − 2)
– Bα = B(4, 2) = 28.3 MeV
Da regra da cadeia, temos também
∂B ∂B
∆B = ∆A + ∆Z
∂A ∂Z
Já sabemos que ∆A = 4, ∆Z = 2. Considera-se A, Z >> 1.
∂B ∂B
∆B = 4 +2
∂A ∂Z
Da fórmula semi-empı́rica,

(N − Z)2 δ
B(Z, N ) = av A − as A2/3 − aC Z(Z − 1)A−1/3 − asym − 1/2
A A
Então
   2
8 Z 2Z
∆B = 4av − as A−1/3 −4ac ZA−1/3 1 − +4asym 1 − +2δA−3/2
3 3A A

Logo
   2
8 −1/3 −1/3 Z 2Z
Q ≈ 28.3−4av + as A +4ac ZA 1− −4asym 1 − +2δA−3/2
3 3A A

75
• Energia disponı́vel para o decaimento:
A−4 4
De A 2 2
Z x → Z−2 Y + 2 He junto com E = mc ⇔ ∆E = ∆mc , temos

Ty + Tα = mX c2 − mY c2 − mα c2 = Q

Logo,
mX c2 = mY c2 + Ty + mα c2 + Tα
mY vY2 mα vα2
Q = TY + Tα = +
2 2
Decaimentos alfa geralmente tem Tα < 9 MeV, podemos usar mecânica
não relativista com segurança: P = mv. Então

PY2 P2
Q= + α
2mY 2mα
Por conservação de momento linear, temos

PY = −Pα

Com as duas equações anteriores temos


Q
Tα =
1 + mα /mY

Mas mα /mY = 4/(A − 4). Com A >> 4, temos 1 + mα /mY = A/(A − 4),
logo
Q
Tα =
A/(A − 4)
Tα = Q(1 − 4/A)

• A energia cinética Tα pode ser medida diretamente com um espectrômetro


magnético, e já sabemos A, então da fórmula acima podemos determinar
o Q.
• Se A ∼ 200, a partı́cula α carrega consigo ∼ 98% do Q, com o fragmento
nuclear muito mais pesado Y levando o resto.
• Para Q ∼ 5 MeV, a energia de recuo do núcleo é TY ∼ 100 keV. Esta
energia é maior que a energia de ligação da rede cristalina.
• Se a transição se der para um estado excitado, há que descontar a energia
do nı́vel
Q = Q0 − Eexc

• Relação de Geiger-Nuttal: relação empı́rica entre o perı́odo de semivida


e a energia cinética da partı́cula α. Ver figura 51.
bZ
log10 (T1/2 ) = a + √

76
Figure 51: .

• Um fator de 2 na energia Q corresponde a um fator de 1024 no perı́odo de


semivida.
• Regras de seleção:
O decaimento α se dá pela força FORTE, que conserva momento angular J
e paridade P . Além disso, os spins do par de protões e do par de neutrões
são ambos 0, logo Jα = lα + sα = lα . Se a reação for X → Y + α, temos
as regras de seleção seguintes
– conservação de J: Jα = JX + JY , ..., |JX − JY |
– conservação de P : PX PY = Pα = (−1)Jα = (−1)lα

10.2 Decaimento Beta


• Partı́cula β: e− ou e+ energético/rápido
• Decaimento β é quando um neutron/próton decai para um próton/neutron,
um elétron/pósitron e antineutrino/neutrino do elétron.
• Decaimento β − : A
ZX
A
→ Z+1 Y + e− + ν e
• Decaimento β + : A
ZX
A
→ Z−1 Y + e+ + νe
• Captura eletrónica: A
ZX + e

→ Z−1 A
Y + νe (quando ocorre decaimento
+
β junto com absorção de um elétron)
• Balanço energético de Qβ − :
Qβ − = [mN (A, Z) − mN (A, Z + 1) − me ]c2
onde mN é a massa do núcleo apenas. A massa atómica (i.e. núcleo +
elétrons) é dada por
Z
X
m(A, Z)c2 = mN (A, Z)c2 + Zme c2 − Bi
i=1

77
Em que Bi = energia de ligação do i-ésimo elétron. Substituindo na
expressão para Q obtemos
Z
X Z+1
X
Qβ − = {[m(A, Z)−Zme ]−[m(A, Z+1)−(Z+1)me ]−me }c2 + Bi − Bi
i=1 i=1

PZ PZ+1
Negligenciando i=1 Bi − i=1 Bi ≈ 0, obtemos

Qβ − = [m(A, Z) − m(A, Z + 1)]c2

O valor Q também corresponde a energia compartilhada pelo elétron e


neutrino:
Qβ − = Te + Eν
Ou seja, o máximo de um corresponde à situação de mı́nimo do outro:

Qβ − = (Te )max = (Eν )max

• Balanço energético de Qβ + :

Qβ + = [mN (A, Z) − mN (A, Z − 1) − me ]c2

Z
X Z+1
X
Qβ + = {[m(A, Z)−Zme ]−[m(A, Z−1)−(Z−1)me ]−me }c2 + Bi − Bi
i=1 i=1
PZ PZ+1
Novamente i=1 Bi − i=1 Bi ≈ 0, pelo que temos

Qβ + = [m(A, Z) − m(A, Z − 1) − 2me ]c2

• Balanço energético de QCE (captura de elétron):


O mesmo raciocı́nio nos dá

QCE = [m(A, Z) − m(A, Z − 1)]c2 − Be

• Avaliação da energia de recuo do


√ núcleo:
2
P2
De T = mv 2 = 2m , temos P = 2mT . Portanto, o momento do núcleo é
p
PN = 2mN TN

Já a partı́cula β terá q


Pβ c = Eβ2 − m2e c4

Mas Eβ = Tβ + me c2 (temos que levar em conta a massa de repouso já


que a β é relativista). Logo
q
Pβ c = (Tβ + me c2 )2 − m2e c4

78
q
Pβ c = (Tβ2 + 2Tβ me c2

Da conservação do momento angular, se Pν ≈ 0, então PN = −Pβ , por-


tanto PN2 = Pβ2 e
1
2mN TN = 2 (Tβ2 + 2Tβ me c2 )
c
Isolando TN , a energia de recuo do núcleo:

Tβ2 me
TN ≈ + Tβ
2mN c2 mN

• Regras de seleção:
Para o decaimento β é requerido que, para o elétron e neutrino, l = 0 (só
aceita). Então j = l + s = s. Elétrons e neutrinos são fermiões ⇔ tem
spin s = ±1/2.
– Se tiverem spins antiparalelos, s = 0, pelo que δj = 0.
– Se tiverem spins paralelos, s = 1, pelo que δj = 0, 1, exceto se
jX = jY = 0 pois nesse caso s = 0 (caso anterior).
Além disso, l = 0 significa que P = (−1)0 = 1, logo não pode mudar a
paridade.
Se a reação for X → Y + α, temos as regras de seleção seguintes

– ∆j = 0, 1
– ∆π (mudança de paridade) = não
• Grau de proibição: Ver figura 52

Figure 52: Graus de proibição

79
10.2.1 Teoria de Fermi do decaimento Beta
• A reação do decaimento β − é n → p + e− + ν ⇔ n + ν → p + e− .
• É uma interação pontual. Ver figura 53.

Figure 53: .

• A interação é descrita por um parâmetro apenas: GF .


• As partı́culas são descritas usando a aproximação mais simples para as
suas funções de onda (i.e. ondas planas)
• A influência do campo eletrostático do núcleo é tomada em consideração
através de uma função de correção F (Z, Ee ), função de Fermi

• Número de desintegrações por intervalo de momento:


dN
= cte. F (Z, Ee )p2e (E − Ee )[(E − Ee )2 − m2ν c4 ]1/2
dpe
Mas mν ≈ 0, então
dN
= cte. F (Z, Ee )p2e (E − Ee )2
dpe

• Energia total do sistema e + ν:

E = Ee + Eν = Te + me + Tν = Q + me (mν ≈ 0)

E − Ee = Q + me − Ee = Q + me − (Te + me ) = Q − Te
dN
= cte. F (Z, Ee )p2e (Q − Te )2
dpe
• Experimentalmente, é mais fácil exprimir a distribuição de eventos em ter-
mos da energia cinética do eletrão. Então, dado que Ee2 = (Te + me c2 )2 =

(pe c)2 = Ee2

80
10.3 Decaimento Gama
• A maioria dos decaimentos α e β deixam o núcleo filho Y num estado
excitado. Estes estados excitados rapidamente emitem um fóton γ, ficando
no estado fundamental. Esta emissão do fóton que é o decaimento gama.
• A desexcitação também pode se dar por conversão interna, que compete
probabilisticamente com o decaimento gama.
• Balanço de energia:
Considere-se um núcleo de massa M em repouso num estado Ei . O núcleo
então decai para o estado Ef . Por conservação de momento, núcleo sofrerá
recoil quando emitir o fóton. Presumimos que este recoil é não-relativista,
tal que a energia cinética é TR = p2R /2M .

Ei = Ef + Eγ + TR (conservação de energia)
0 = p⃗R + p⃗γ (conservação de momento)
Logo,
p2R
∆E = Ei − Ef = Eγ +
2M
2 2
Mas da cons. de momento temos pR = pγ , e sabemos que pγ = cEγ . Logo,

Eγ2
∆E = Eγ +
2M c2
Deixamos na forma de eq. quadrática ax2 + bx + c = 0:
1
E 2 + Eγ − ∆E = 0
2M c2 γ
Resolvendo, " r #
2 2∆E
Eγ = M c −1 ± 1+
M c2

Mas ∆E(∼ MeV) << M c2 (∼ A × 103 MeV), pelo que podemos fazer
uma boa aproximação com os primeiros 3 termos da expansão em série de
Taylor:

(1 + x)α ≈ 1 + αxα−1 + ((1 + x)α−1 + α(α − 1)(1 + x)α−2 )

No nosso caso, ficamos com


(∆E)2
Eγ ≈ ∆E −
2M c2

• Dipolos elétrico e magnético: ver figura 54


– Dipolo elétrico: consiste de cargas iguais e opostas q e −q, separadas
por uma distância fixa z ⇒ momento dipolar elétrico d = qz.

81
– Dipolo magnético: consiste de um loop circular de corrente i e área
A ⇒ momento dipolar magnético µ = iA.
• Nós podemos produzir campos de radiação EM variando os momentos
dipolares; por exemplo,

– podemos fazer as cargas oscilarem pelo eixo z, tal que d(t) = qzcos(ωt).
– podemos variar a corrente tal que µ(t) = iA cos(ωt)
• Os campos dipolares E e M tem paridades opostas: considere a trans-
formação de paridade r → −r. Olhando para a figura 54, vê-se que

– para o dipolo elétrico, E(r) = E(−r), B(r) = −B(−r).


– para o dipolo magnético, E(r) = −E(−r), B(r) = B(−r).
Logo, a paridade dos campos radiativos é diferente:
– o campo dipolar elétrico tem paridade ı́mpar: π(EL) = (−1)L
– o campo dipolar magnético tem paridade par: π(M L) = (−1)L+1
• Ordem do multipolo: 2L , tal que L = 1 para o dipolo, L = 2 para o
quadrupolo, L = 3 para octopolo etc.
• Estimativas de Weisskopf:

• Regras de Seleção:

– |ji − jf | ≤ l ≤ ji + jf
– l ̸= 0
– ∆π = não: elétrico par, magnético ı́mpar
– ∆π = sim: elétrico ı́mpar, magnético par

82
Figure 54: .

83

You might also like