theonp_ch2
theonp_ch2
Angular momenta and the coupling of angular momenta play a significant role in theoret-
ical nuclear physics. Therefore, the quantum mechanical description of angular momenta
will be briefly reviewed here.
The commutators between the components satisfy the so-called angular momentum alge-
angular
bra momentum
[ jˆx , jˆy ] = i jˆz and cyclically . (2·2) algebra
ˆ
This is the only information about the properties of the operator ⃗j that we have – and the
only information we need to determine the structure of the eigenbasis.
From the commutation relations, it is immediately apparent that no simultaneous
ˆ
eigenbasis of two Cartesian components of ⃗j can be formed. Nevertheless, it is possible to
form an operator from the components that commutes with each of the components. This
operator is the square of the angular momentum operator
⃗jˆ 2 = jˆ2 + jˆ2 + jˆ2 . (2·3)
x y z
Using the angular momentum algebra (2·2), it can be easily shown that
ˆ
[ ⃗j 2 , jˆz ] = 0 (2·4)
ˆ
holds. Thus, there exists a simultaneous eigenbasis of ⃗j 2 and one of the Cartesian compo- simultaneous
nents, conventionally chosen to be jˆz . The eigenvalue equations that define the simulta- eigenbasis
Here we use a very compact notation where the eigenstates are identified merely by the
formula symbols of the respective eigenvalues.
The eigenvalues j and m cannot take arbitrary values but are quantized. From the
angular momentum commutation relations (2·2), the possible eigenvalues, i.e., the quan-
tization rules, can be derived. For this purpose, ladder operators jˆ± = jˆx ± i jˆy are usually
defined, and their effect on the eigenvectors | jm ⟩ is examined. It follows that j can only
take integer or half-integer values
j = 0, 21 , 1, 32 , ... . (2·6)
For a given j, the quantum number m associated with the z-component (magnetic quantum
number, projection quantum number) can take only (2j + 1) different values
Using the canonical commutation relations between position and momentum operators
[ x̂i , p̂ j ] = i δi j (2·9)
it can be immediately shown ({ exercise), that the orbital angular momentum operator
obeys the angular momentum algebra
Thus, the results of the previous section can be directly applied to the orbital angular
momentum operator. In particular, the eigenvalue equations that define the simultaneous
eigenbasis have the same structure as in the general case:
In addition to the commutation relations, the definition (2·8) provides another con-
straint on the possible quantum numbers. To derive this, we need to switch to the position
representation and examine the structure of the eigenfunctions of the orbital angular mo-
mentum operator. After some calculations, the following position representations of the
ˆ
operators ⃗l 2 and l̂z are obtained from the definition of the orbital angular momentum
operator:
ˆ 1 ∂2 1 ∂ ∂ i
x | ⃗l 2 | ψ ⟩ = −
h
⟨⃗ + sin θ ⟨⃗
x|ψ⟩
sin2 θ ∂ϕ2 sin θ ∂θ ∂θ
(2·12)
∂
⟨⃗
x | l̂z | ψ ⟩ = −i ⟨⃗
x|ψ⟩ .
∂ϕ
This allows the eigenvalue equations to be transferred to the position representation. The
solutions to the resulting differential equations are the spherical harmonics, which thus spherical
harmonics
form the position representation of the abstract eigenstates | lml ⟩:
Due to the uniqueness of the spherical harmonics, only integer values of l and ml are
allowed:
l = 0, 1, 2, ...
(2·14)
ml = −l, −l + 1, ..., l − 1, l .
This is particularly evident from the ϕ-dependence of the spherical harmonics, which is
given by exp(i ml ϕ). Only integer ml lead to a unique value at ϕ = 0 and 2π.
ˆ
For later use, we note that the position representation of ⃗l 2 directly enters into the
position representation of the operator ⃗pˆ2 :
1 ∂2 1 ˆ2
ˆ
x|⃗
⟨⃗ p |ψ⟩ = −
2
r⟨ ⃗
x|ψ⟩ − 2⟨⃗ ⃗
x|l |ψ⟩ . (2·15)
r ∂r2 r
2·1·3 Spin
Spin, unlike orbital angular momentum, does not act in position space but forms an
independent spin space. A complete basis of the spin space is defined by the above eigen-
states. Elementary particles are assigned different values of the spin quantum number s:
protons, neutrons, and electrons have spin s = 1/2, pions have spin s = 0, and photons
have spin s = 1. The value of the magnetic spin quantum number ms can take all permissi-
ble values −s, −s + 1, ..., s − 1, s, so that the spin space has the dimension (2s + 1) for a given
s.
Spin s = 1/2
Of particular interest for our applications is the case s = 1/2. The basis in the spin Hilbert
space Hspin is two-dimensional and is spanned by the two eigenstates for ms = ±1/2. The
following designations for the two basis states are commonly used:
| 21 , + 12 ⟩ = | ↑ ⟩ ,
(2·17)
| 21 , − 12 ⟩ = | ↓ ⟩ .
Since the spin-1/2 space is so small, there are some special relations for the spin oper-
ator itself. For example, the following holds [see Bransden&Joachain, Sec. 6.8]:
and
i
ŝx ŝ y = −ŝ y ŝx =
ŝz and cyclically . (2·19)
2
From these relations, it directly follows that any product of spin operators can be reduced
to a single operator or a number. Furthermore, it follows that any operator in the spin
space can be represented as a linear superposition of {1̂, ŝx , ŝ y , ŝz }. This set of operators
thus forms a complete, linearly independent basis of operators in the spin-1/2 space. Note:
This only applies to the spin Hilbert space for s = 1/2!
To eliminate the ubiquitous factors of 1/2 when working with the spin operator, a
reduced spin operator is often used in nuclear physics:
In the following, both versions will be used, with ⃗sˆ always denoting the full spin operator,
and ⃗σˆ being reserved for the reduced spin operator.
Since the basis for s = 1/2 is only two-dimensional, an explicit basis representation of
the components of the spin operator can easily be given. Each component of the abstract
operator is thus associated with a 2 × 2 matrix in the basis representation:
⟨ ↑ | σ̂i | ↑ ⟩ ⟨ ↑ | σ̂i | ↓ ⟩
σi = . (2·21)
⟨ ↓ | σ̂ | ↑ ⟩ ⟨ ↓ | σ̂ | ↓ ⟩
i i
Pauli spin These are the so-called Pauli spin matrices. The matrix elements can easily be calculated
matrices
explicitly. For σz this is trivial, as we are in the eigenbasis of the operator; for σx and σ y
the matrix elements can be derived using the ladder operators σ̂± = σ̂x ± i σ̂ y . The result
is:
0 1 0 −i 1 0
σx = , σ y = , σz = . (2·22)
1 0 i 0 0 −1
It is important to distinguish between the matrices σi and the operators σ̂i – the Pauli spin
matrices are not abstract operators acting on states of the spin Hilbert space, but merely
complex-valued matrices. All relations for the abstract spin operator for s = 1/2 transfer
directly to the Pauli spin matrices.
spinor rep- In (older) literature, one often finds the so-called spinor representation for the state of
resentation
a particle in spin space. This is merely the basis representation of a general spin state,
which can be written as a superposition of the two basis states | ↑ ⟩ and | ↓ ⟩:
| χ ⟩ = c↑ | ↑ ⟩ + c↓ | ↓ ⟩ . (2·23)
If the basis states are identified analogously to the spin matrices with two-component
complex-valued vectors, a spinor results, which contains the amplitudes for the two basis
components as entries
⟨ ↑ | χ ⟩ c↑
⃗ =
χ = .
(2·24)
⟨ ↓ | χ ⟩ c
↓
With spin matrices and spinors, one can calculate in the same way as with spin operators
and spin states. For reasons of consistency, we will generally use the operator formulation.
2·1·4 Isospin
Isospin (isotopic spin) is an invention of Heisenberg, which he discussed in his first pub-
lication on nuclear physics in 1932 [Z. Phys. 77, 1 (1932)]. It is a purely formal quan-
tum number, an auxiliary technical quantity, unlike spin and orbital angular momentum.
Heisenberg describes proton and neutron as two different intrinsic states of a single par-
ticle, the nucleon. Since proton and neutron have approximately the same mass, this
approach seems at least approximately justified.
For the formal definition of an isospin quantum number in quantum mechanics, we
need an operator whose eigenbasis is exactly two-dimensional. The two eigenstates can
then be identified with proton and neutron. As just discussed, the spin operator in the
subspace for s = 1/2 has a two-dimensional eigenbasis. Therefore, in analogy, an isospin
ˆ
operator ⃗t = (t̂1 , t̂2 , t̂3 ) is introduced, which formally fulfills the angular momentum com-
mutation relations
[ t̂1 , t̂2 ] = i t̂3 and cyclically . (2·25)
Note that we index the cartesian components with {1, 2, 3} and not with {x, y, z}. This is to
indicate that the isospin operator is defined in a purely formal three-dimensional space,
which is completely independent of the three-dimensional spatial space. This distinction
is necessary, for example, when it comes to coupling angular momenta or the transfor-
mation behavior of states under spatial rotations. It makes no sense to define a formal
quantum number to distinguish between proton and neutron that would describe a tran-
sition between proton and neutron as a consequence of a spatial rotation – that would
be the case if the components of the isospin operator were defined in three-dimensional
spatial space. Similarly, it makes no sense to couple the formal isospin quantum number
with orbital angular momentum, for example. All these points differentiate isospin from
spin – the latter transforms under spatial rotations and can be coupled with other angular
momenta.
The eigenvalue equations are analogous to the general angular momentum equations:
with the possible quantum numbers t = 0, 21 , 1, ... and mt = −t, −t + 1, ..., t. For the de-
scription of the nucleon, the case t = 1/2 is relevant, where the desired two-dimensional
eigenbasis is spanned by
| 12 , + 12 ⟩ = | p ⟩ ↔ proton
(2·27)
| 21 , − 12 ⟩ = | n ⟩ ↔ neutron
| 1, +1 ⟩ = | π+ ⟩ , | 1, 0 ⟩ = | π0 ⟩ , | 1, −1 ⟩ = | π− ⟩ . (2·29)
The four ∆ baryons ∆++ , ∆+ , ∆0 and ∆− form an isospin quartet with t = 3/2:
| 32 , + 32 ⟩ = | ∆++ ⟩ , | 32 , + 21 ⟩ = | ∆+ ⟩ ,
(2·30)
| 32 , − 12 ⟩ = | ∆0 ⟩ , | 32 , − 32 ⟩ = | ∆− ⟩ .
It turns out that the isospin formalism has more physical content than originally expected.
It systematically appears as a quantum number linked with charge in the characterization
of elementary particles.
For a complete description of the state of a nucleon, the three spaces are combined using
the Kronecker product (tensor product)
The dimension of the resulting Hilbert space corresponds to the product of the dimensions
of the individual spaces.
The simplest single-particle state | α ⟩ ∈ H1 is, according to the structure of the Hilbert
space, a Kronecker product of a state | ϕ ⟩ in the spatial space Hspatial , a state | χ ⟩ in the
spin space Hspin , and a state | ξ ⟩ in the isospin space Hisospin :
These product states form only the simplest class of states in the combined Hilbert space product
states
H1 – not every single-particle state is a product state. However, every single-particle state
can be represented as a superposition of product states.
A complete basis in the product space H1 can be constructed directly from the bases
of the subspaces. For the spin and isospin space of the nucleon, the bases are trivially
given by the two eigenstates { | ↑ ⟩, | ↓ ⟩} and { | p ⟩, | n ⟩}, respectively. For the spatial space,
various bases can be considered depending on the problem. For example:
| ⃗k ⟩ = | kx k y kz ⟩ pˆ or the kinetic
Eigenstates of the momentum operator ⃗
energy (plane waves)
| nx n y nz ⟩ Eigenstates of the (deformed) harmonic oscillator
characterized by Cartesian quantum numbers
| nlml ⟩ Eigenstates of a spherically symmetric potential char-
acterized by radial and angular momentum quantum
numbers
Initially, we use ν as a generic collective index for the quantum numbers of any basis in
Hspatial .
The completeness of the basis allows the decomposition of the identity operator on H1
in the product basis
X
1̂ = ( | ν ⟩ ⊗ | ms ⟩ ⊗ | mt ⟩)(⟨ ν | ⊗ ⟨ ms | ⊗ ⟨ mt | )
ν,ms ,mt
X (2·33)
= | νms mt ⟩⟨ νms mt | .
ν,ms ,mt
In the second line, we use a shorthand notation for the product states
| νms mt ⟩ ≡ | ν ⟩ ⊗ | ms ⟩ ⊗ | mt ⟩ , (2·34)
which no longer explicitly reveals the structure of the individual subspaces. This notation
is very convenient, but one must keep in mind the assignment of quantum numbers to the
various subspaces. Using the decomposition, the basis expansion of any state | α ⟩ can be
easily formulated:
X
|α⟩ = | νms mt ⟩⟨ νms mt | | α ⟩
ν,m ,mt
Xs (2·35)
= Cνms mt | νms mt ⟩ ,
ν,ms ,mt
| α ⟩ ⊗ (x | β ⟩) ⊗ | γ ⟩ = x( | α ⟩ ⊗ | β ⟩ ⊗ | γ ⟩) , (2·38)
| α ⟩ ⊗ ( | β ⟩ + | β′ ⟩) ⊗ | γ ⟩ = ( | α ⟩ ⊗ | β ⟩ ⊗ | γ ⟩) + ( | α ⟩ ⊗ | β′ ⟩ ⊗ | γ ⟩) . (2·39)
The scalar product between product states reduces to the product of the scalar products
of the partial states from the respective subspaces
⟨ ψ | ψ′ ⟩ = ⟨ α | ⊗ ⟨ β | ⊗ ⟨ γ | | α′ ⟩ ⊗ | β′ ⟩ ⊗ | γ′ ⟩
(2·40)
= ⟨ α | α′ ⟩ ⟨ β | β′ ⟩ ⟨ γ | γ′ ⟩ .
The same applies to the dyadic product | ψ ⟩⟨ ψ′ | . It is important to note: only the scalar
product between states of the same space makes sense. The scalar product between
states from different spaces (e.g., spatial and spin space) is mathematically undefined.
The Kronecker product is also necessary to define operators on the product space.
Suppose the operator  is defined on Ha , B̂ on Hb , and Ĉ on Hc , then
X̂ = Â ⊗ B̂ ⊗ Ĉ . (2·41)
describes an operator on the product space. Again, the order of the factors is crucial.
For example, it makes no sense to apply a spin operator to the spatial space. Here are
the essential computational rules for the Kronecker product with operators:
The matrix element of the operator X̂ between product states decomposes into a prod-
uct of matrix elements for the individual subspaces:
⟨ ψ | X̂ | ψ′ ⟩ = ⟨ α | Â | α′ ⟩⟨ β | B̂ | β′ ⟩⟨ γ | Ĉ | γ′ ⟩ . (2·45)
Finally, a warning: often, a sloppy notation for operators on product spaces is used.
Consider, for example, the spin operator ⃗sˆ on a space with the structure (2·31). The
correct transfer of the spin operator to the product space forces the following Kronecker
product:
⃗sˆ → 1̂ ⊗ ⃗sˆ ⊗ 1̂ , (2·46)
i.e., the operator acts like the identity operator on the first and third subspace. This
notation is quite cumbersome and is therefore used only when the operators cannot be
intuitively assigned to the respective subspaces.
If the individual subspaces are identical and thus an intuitive assignment is not possi-
ble, then the operator symbol is followed in parentheses by the number of the subspace
in which it acts in a compact notation. In a space consisting of several identical sub-
spaces, we write for a one-body operator â that acts only on the i-th subspace:
Âi = 1̂ ⊗ · · · ⊗ â ⊗ · · · ⊗ 1̂ . (2·47)
i-th subspace
Note that â is an operator acting on a single-particle Hilbert space, while Âi is acting on
the many-particle space. This notation will become important in the following discus-
sion of many-particle Hilbert spaces for describing systems of identical particles. This
notation can also be extended directly to operators acting simultaneously in multiple
subspaces – the affected subspaces are then listed together in the parentheses.
So far, we have only discussed the Hilbert space H1 of a single nucleon. Now, the task is
to construct the Hilbert space to describe an entire nucleus, i.e., a system of A indistin-
guishable nucleons.
Intuitively, each of the A nucleons brings its degrees of freedom, which are summarized
in the single-particle space H1 , into the many-particle space. Therefore, the A-particle
space should be expressible as a ”combination” of the A single-particle spaces in the sense
of a Kronecker product:
HA = H1 ⊗ H1 ⊗ · · · ⊗ H1 . (2·48)
product On this product space, product states can be defined, consisting of Kronecker products of
states
A single-particle states | α1 ⟩, | α2 ⟩, ..., | αA ⟩:
| α1 , α2 , . . . , αA ⟩ = | α1 ⟩ ⊗ | α2 ⟩ ⊗ · · · ⊗ | αA ⟩ . (2·49)
Here, α is used as a generic index for all quantum numbers of the nucleon, e.g., α =
n, l, ml , ms , mt . The left side shows a commonly used shorthand notation for the product
states, where the order of the single-particle indices in the ket symbol is essential. The
totality of product states from all possible combinations of the states | α ⟩ of a complete
single-particle basis forms a basis of the product space HA .
Although product states have a conveniently simple structure, they contradict a central
physical property of many microscopic many-particle systems: the indistinguishability of
indistinguish- particles. The individual protons and neutrons of a nucleus cannot be distinguished based
ability of
particles on any immutable property. Hence, statements like ”Nucleon number 5 is in the state α5 ”
make no sense – one cannot assign a unique number to any of the nucleons. However,
the product state (2·49) does exactly that, establishing a one-to-one correspondence be-
tween single-particle spaces (the particles) and single-particle states. Therefore, product
states are not allowed for indistinguishable particles – they apply only to distinguishable
particles.
The consequences of particle indistinguishability can be formulated as follows: Every
observable, e.g., scalar products, expectation values, or matrix elements, must be inde-
pendent of any formal order or assignment of single-particle states. A simple observable is
the probability density, which follows from the squared magnitude of the wave function,
e.g., in position representation. Considering a general A-particle state | ψ ⟩, the squared
magnitude of the corresponding wave function must satisfy2 :
2 2
⟨⃗
x1 , ⃗
x2 , . . . , ⃗
xA | ψ ⟩ = ⟨ ⃗
x1 , ⃗
x2 , . . . , ⃗
xA | T̂i j | ψ ⟩ . (2·50)
transposition Here, T̂i j is the so-called transposition operator, which swaps the single-particle states in
operator the single-particle spaces i and j. Formally, this can be summarized in a criterion for
allowed many-particle states: The indistinguishability of particles implies that every phys-
ically meaningful many-particle state satisfies the following condition:
T̂i j | ψ ⟩ = ± | ψ ⟩ ∀i, j . (2·51)
2
Spin and isospin degrees of freedom are not considered here for clarity.
That is, many-particle states must be eigenstates of the transposition operator with eigen-
values ±1.
A similar analysis can be performed for the expectation value or the matrix element of
any operator Â. Again, this observable must be independent of the assignment of single-
particle quantum numbers in the states. This must hold for arbitrary states, regardless of
their transposition symmetry3 . Therefore, the operator of any observable must be invari-
ant under transpositions, i.e.,
or equivalently
[ T̂i j , Â ] = 0 ∀i, j . (2·53)
A trivial consequence of this property is that there should be no operators that act only on
individual subspaces. Every meaningful operator must act on all single-particle spaces in
the same way in the sense of indistinguishability.
Returning to the many-particle states, three different classes of many-particle states can
be identified based on their behavior under transpositions. Accordingly, the many-particle
Hilbert space splits into three subspaces:
Only the states with defined transposition symmetry describe physical systems; states from
HAindef are not realized for systems of identical particles.
There are significant differences between systems described by symmetric and antisym-
metric states. In the quantum statistical description of ideal gases, it is found that both
classes of states result in different distribution functions f (E) or f (k). Symmetric states
lead to a Bose-Einstein distribution, while antisymmetric states result in a Fermi-Dirac dis-
tribution. Accordingly, particles described by symmetric many-particle states are called
bosons. Particles associated with antisymmetric many-particle states are called fermions. bosons
3
Operators are usually defined for application on arbitrary spaces, e.g., the transposition operator, which
can be applied to all kinds of many-particle states. fermions
The question remains by what criteria individual elementary particles, such as nucleons
or electrons, can be classified into one of these categories. From the theoretical side, this
spin-
statistics question is answered by the spin-statistics theorem. It states that particles with integer spin
theorem are bosons and particles with half-integer spin are fermions. The proof of this theorem
requires the full apparatus of relativistic quantum field theory, which we cannot delve into
here. Essentially, the spin-statistics theorem arises from the requirement of microcausality.
Empirically, the connection between spin and quantum statistics has been confirmed many
times.
In the following table, these relationships are summarized:
It is remarkable that not only elementary particles but also complex, composite objects,
such as entire atoms, obey the spin-statistics theorem. The key factor is the energy scales
at which one wishes to describe the many-particle system. At very low temperatures,
4 He and 3 He atoms behave like inert bosons and fermions, respectively, where the total
angular momentum of the nucleus and the shell electrons takes the role of spin. When the
energy is increased so that the substructure of the atoms is resolved—at the latest when
the energy is sufficient to excite the atom—this effective picture loses its validity.
In recent years, this has been impressively demonstrated with ultracold atomic gases
in magnetic traps. At temperatures in the nano-Kelvin range, even heavier atoms, such as
23 Na or 87 Rb, behave like perfect bosons and exhibit dramatic phenomena such as Bose-
Einstein condensation.
The simple product states (2·49) are not physically meaningful many-particle states for a
system of identical particles. Nevertheless, they form a basis in the product space HA ,
which contains the space of symmetric and antisymmetric states, HAs and HAa , as sub-
spaces. It must, therefore, be possible to generate a basis for HAs and HAa from the basis
of product states by projecting onto the respective subspaces.
The structure of this projection can be explained intuitively: To create a symmetric
state from product states, we must superpose the product states for all possible arrange-
ments of the single-particle indices. The resulting superposition state is automatically
symmetric with respect to transposition. Formally, this recipe can be implemented with
symmetrizationthe following projection operator, the so-called symmetrization operator:
operator
1 X
S= P̂π . (2·55)
A! π
Here, the symbol π denotes a permutation, i.e., a reordering scheme for the indices
{1, 2, . . . , A} → {π(1), π(2), . . . , π(A)}. The summation runs over all A! possible permuta-
tions of the indices. The permutation operator P̂π rearranges the single-particle indices of
a many-particle state according to the given permutation π. Applied to a product state, it
yields:
P̂π | α1 , α2 , . . . αA ⟩ = | απ(1) , απ(2) , . . . απ(A) ⟩ . (2·56)
Thus, we can easily understand the operation of the operator S: Applied to a product
state, it generates a superposition of product states for all possible permutations of the
single-particle states.
anti-
Analogously, an antisymmetrization operator can be constructed. Compared to the sym- symmetrization
metrizer, we only need to choose the sign for each state in the superposition appropriately. operator
For this, we use the fact that any permutation can be represented as a sequence of trans-
positions, with the number of elementary transpositions defining the signature sgn(π) of
the permutation:
sgn(π) = (−1)number of transpositions in π . (2·57)
This defines the necessary sign of the individual summands. Thus, we can directly write
the definition of the antisymmetrization operator:
1 X
A= sgn(π) P̂π . (2·58)
A! π
Symmetrizer and antisymmetrizer are projection operators, i.e., they are Hermitian and projection
operators
idempotent ({ exercise):
S† = S , S2 = S , (2·59)
A† = A , A2 = A . (2·60)
SA = 0 . (2·61)
A direct consequence of this is that scalar products or matrix elements between sym-
metrized and antisymmetrized states vanish—this is occasionally referred to as a superse-
lection rule.
In the literature, there are also definitions of the symmetrization and antisymmetriza-
tion operators that do not include the prefactor 1/A!. The projector property, which will
be useful later, is thus lost.
We will now focus on the antisymmetric Hilbert space HAa , which is relevant for the nu-
clear many-particle problem.
{ | α1 , α2 , α3 ⟩a , | α2 , α1 , α3 ⟩a , | α1 , α3 , α2 ⟩a ,
(2·63)
| α3 , α2 , α1 ⟩a , | α2 , α3 , α1 ⟩a , | α3 , α1 , α2 ⟩a } ,
as can be immediately seen from the explicit representation as a sum over all permutations
({ exercise). Only one can be used as a basis state for the antisymmetric subspace. In
other words: Each set of single-particle quantum numbers {α1 , α2 , . . . , αA } is associated
with only one antisymmetrized basis state – regardless of the order. Since there are A!
permutations in total, the basis of the antisymmetric subspace HAa is smaller by at least a
factor of 1/A! than that of the full product space HA . This point is essential when it comes
to utilizing the completeness of the antisymmetrized product basis.
The antisymmetrized product states automatically form an orthonormal set, provided
the single-particle states used are orthonormal,
Furthermore, the completeness allows the following decomposition of the identity opera-
tor on the antisymmetric subspace HAa :
X
1̂ = | α1 , α2 , . . . , αA ⟩aa ⟨ α1 , α2 , . . . , αA |
α1 <α2 <···<αA
1 X (2·65)
= | α1 , α2 , . . . , αA ⟩aa ⟨ α1 , α2 , . . . , αA | .
A! α ,α ,...,α
1 2 A
Here, the problem of linear dependence discussed above is immediately evident. In the
summation, one must either restrict to a sorting of the single-particle quantum numbers,
formally indicated by α1 < α2 < · · · < αA , or correct for the multiple counting of the
individual basis states by the factor 1/A!.
A direct consequence of the antisymmetry of the many-particle state is the Pauli prin-
Pauli ciple. It is not possible to construct an antisymmetric many-particle state if two particles
principle
occupy the same single-particle state. The antisymmetrizer automatically maps a product
state that contains two identical single-particle indices to zero ({ exercise).
The special structure of the antisymmetrizer corresponds to the formal definition of
determinants, hence antisymmetrized product states are also called Slater determinants. Slater deter-
minants
The position representation of an antisymmetrized product state shows the determinant
structure directly. Neglecting the spin and isospin degrees of freedom, we have
α1 (⃗
x1 ) α1 (⃗
x2 ) . . . α1 (⃗
xA )
1 α2 (⃗x1 ) α2 (⃗x2 ) . . . α2 (⃗xA )
⟨⃗
x1 , ⃗
x2 . . . , ⃗
xA | α1 , α2 , . . . , αA ⟩a = √ .. .. .. .. , (2·66)
A! . . . .
αA (⃗
x1 ) αA (⃗
x2 ) . . . αA (⃗
xA )
x) = ⟨ ⃗
where αi (⃗ x | αi ⟩ denotes the position representation of the single-particle states
(single-particle wave function).
Of great practical importance in atomic and nuclear physics is the quantum mechanical
addition of angular momenta, the so-called angular momentum coupling. We will discuss
some variants here in preparation for further applications.
Before turning to the many-particle system, let us consider the coupling of angular mo-
menta in the Hilbert space of a nucleon. As discussed in Section 2·2·1, the single-particle
Hilbert space has the following structure:
A possible basis in this combined space consists of product states formed from basis states
of the subspaces:
| nlml , 12 ms , 12 mt ⟩ ≡ | nlml ⟩ ⊗ | 21 ms ⟩ ⊗ | 12 mt ⟩ . (2·68)
uncoupled We will refer to this basis as the uncoupled basis. Since the radial quantum number n
basis
and the isospin part of the state are not further relevant, we will restrict ourselves to
the quantum numbers in the angular-spin space with the shorthand notation | lml , sms ⟩.
ˆ
These states are simultaneous eigenstates of the operators ⃗l 2 , l̂z , ⃗sˆ 2 , and ŝz and satisfy the
following eigenvalue equations:
total There is another way to construct a basis in the angular-spin space. For this, we define
ˆ
angular a new total angular momentum operator ⃗j as the sum of the orbital angular momentum
momentum
operator and spin operators in the single-particle system – we couple orbital angular momentum
and spin to a total angular momentum:
This is an abbreviated notation that does not show the assignment of the operators to the
individual subspaces. The detailed definition of the total angular momentum operator in
a space of structure (2·67) is
i.e., the total angular momentum operator acts in the space and spin space.
ˆ
This operator ⃗j satisfies the usual angular momentum commutator relations (2·2), as
can be immediately derived from the commutators for orbital angular momentum and
spin operators:
ˆ
[ ⃗j 2 , jˆz ] = 0 . (2·72)
In addition, the following ”mixed” commutator relations arise directly from the definition
({ exercise):
ˆ ˆ ˆ ˆ
[ ⃗j 2 , ⃗l 2 ] = [ ⃗j 2 , ⃗sˆ 2 ] = [ ⃗l 2 , ⃗sˆ 2 ] = 0 , (2·73)
ˆ
[ jˆz , ⃗l 2 ] = [ jˆz , ⃗sˆ 2 ] = 0 . (2·74)
From these commutator relations, we can identify a second set of operators that com-
mute with each other and thus have a simultaneous eigenbasis. These are the four opera-
ˆ ˆ
tors ⃗l 2 , ⃗sˆ 2 , ⃗j 2 , and jˆz . The simultaneous eigenstates define the so-called coupled basis: coupled
basis
We have introduced a special bracket notation for the states of the coupled basis. This
indicates the coupling scheme used. The symbol ,,(ls)jm” can be read as: ,,l and s coupled
to j and m”.
Both the coupled and the uncoupled states form a complete basis in the angular-spin
space, i.e., we can represent the identity operator in this subspace in both bases:
X
1̂ = | lml , sms ⟩⟨ lml , sms | ,
l,ml ,s,ms
X (2·77)
1̂ = | (ls)jm ⟩⟨ (ls)jm | .
l,s,j,m
This allows us to express the states of one basis in terms of the other. For the development
of the coupled states in the uncoupled basis, we have:
X
| (ls) jm ⟩ = | l′ ml , s′ ms ⟩⟨ l′ ml , s′ ms | (ls) jm ⟩
l′ ,ml ,s′ ,ms
X (2·78)
= | lml , sms ⟩⟨ lml , sms | (ls)jm ⟩ ,
ml ,ms
where we have used the orthogonality of the basis states with respect to l and s, i.e.,
⟨ l′ ml , s′ ms | (ls)jm ⟩ ∝ δl,l′ δs,s′ . The scalar product ⟨ lml , sms | (ls) jm ⟩ provides the coefficients
for the unitary transformation between the coupled and uncoupled bases – these are also
Clebsch-
referred to as Clebsch-Gordan coefficients. We define the following symbolic notation for Gordan
the CG coefficients: coefficients
l s j
c = ⟨ lml , sms | (ls) jm ⟩ . (2·79)
ml ms m
The unitary transformation between the coupled and uncoupled bases is thus given by
X
l s j
| (ls) jm ⟩ = c | lml , sms ⟩ (2·80)
m ,m
m l ms m
l s
and conversely
X
l s j
| lml , sms ⟩ = c | (ls)jm ⟩ , (2·81)
ml ms m
j,m
where we have taken advantage of the fact that the CG coefficients can be chosen to be
real. The numerical values of the CG coefficients are tabulated, an example of such a table
is shown in Figure 2·1. For numerical calculations there are libraries for any programming
language that compute CG coefficients efficiently — no need to worry about the tables.
A summary of the most important properties of the CG coefficients is provided in the
following supplement. It should be emphasized that everything we have illustrated here
using the example of spin-orbit coupling in the single-particle system also applies to the
coupling of other angular momenta. Some applications will be discussed in the following
sections.
s j′
X
l s j l
δ δ
j,j′ m,m′ = c c . (2·84)
m ,m
ml ms m ml ms m′
l s
Figure 35.1: The sign convention is that of Wigner (Group Theory, Academic Press, New York, 1959), also used by Condon and Shortley (The
Theory of Atomic Spectra, Cambridge Univ. Press, New York, 1953), Rose (Elementary Theory of Angular Momentum, Wiley, New York, 1957),
and Cohen (Tables of the Clebsch-Gordan Coefficients, North American Rockwell Science Center, Thousand Oaks, Calif., 1974). The coefficients
here have been calculated using computer programs written independently by Cohen and at LBNL.
F IGURE 2·1 : Schematic compilation of the Clebsch-Gordan coefficients. The individual blocks
show all non-vanishing CG coefficients for given values of the uncoupled angular momenta (l and
s in our example). [Particle Data Group]
• The CG coefficients can be explicitly calculated using the following formula [Ed-
monds 1960, Eq. (3.6.11)]:
s
l s j
(2j + 1)(l + s − j)!(l − s + j)!(−l + s + j)!
c = δm,ml +ms
ml ms m (l + s + j + 1)!
p
× (l + ml )!(l − ml )!(s + ms )!(s − ms )!(j + m)!(j − m)!
X (−1)z
× ,
z
z!(l + s − j − z)!(l − ml − z)!(s + ms − z)!(j − s + ml + z)!(j − l − ms + z)!
(2·89)
After coupling the spin and orbital angular momentum of a single nucleon, we now turn
to the two-particle system. There are many possibilities to couple the various angular
momenta of two nucleons.
First, we want to couple the spins of two nucleons to form a total spin. To simplify, we
consider only the spin Hilbert spaces of the two nucleons and combine them into the
total spin
⃗ˆ is defined on this space as
product space Hspin ⊗ Hspin . The operator of the total spin S
The total spin operator also satisfies the usual angular momentum commutation rela-
tions. Again, we have two sets of four operators each that commute with each other and
therefore have a simultaneous eigenbasis. The uncoupled basis is given by product states
| s1 ms1 , s2 ms2 ⟩ = | s1 ms1 ⟩ ⊗ | s2 ms2 ⟩, formed from the spin eigenstates in the single-particle
space, and satisfies the eigenvalue relations for ⃗sˆ 21 , ŝz,1 , ⃗sˆ 22 , and ŝz,2 .
The coupled basis consists of the eigenstates for ⃗sˆ1 2 , ⃗sˆ2 2 , S ⃗ˆ 2 , and Ŝz . The following
eigenvalue relations hold:
The states of the coupled basis are obtained from the uncoupled basis through transfor-
mation with the CG coefficients:
X
s s2 S
| (s1 s2 )SMS ⟩ = c 1 | s1 ms1 , s2 ms2 ⟩ . (2·92)
m ,m
ms1 ms2 MS
s1 s2
For us, the spin coupling for spin s = 1/2 is particularly important. In this case, the spin s = 1/2
two-particle spin space is four-dimensional. The uncoupled basis consists of the states:
| ↑, ↑ ⟩ = | 12 , + 21 ⟩ ⊗ | 12 , + 12 ⟩ ,
| ↑, ↓ ⟩ = | 21 , + 21 ⟩ ⊗ | 12 , − 12 ⟩ ,
(2·93)
| ↓, ↑ ⟩ = | 21 , − 21 ⟩ ⊗ | 12 , + 12 ⟩ ,
| ↓, ↓ ⟩ = | 21 , − 21 ⟩ ⊗ | 12 , − 12 ⟩ .
The coupled basis must also consist of four states. From the properties of the CG
coefficients (triangle inequality), it immediately follows that only total spins S = 0 and
S = 1 can be generated through the coupling. From the quantization rules for angular
momentum, it further follows that for S = 0 only MS = 0 is possible, and for S = 1 the
three values MS = −1, 0, 1 are possible. Using the numerical values of the CG coefficients,
the explicit representation of the four states of the coupled basis can be derived. For the
one state with S = 0, the so-called singlet state, we have singlet state
| 0, 0 ⟩ = √1 ( | ↑, ↓ ⟩ − | ↓, ↑ ⟩) . (2·94)
2
For the three states with S = 1, the so-called triplet states, we get: triplet states
| 1, +1 ⟩ = | ↑, ↑ ⟩ ,
| 1, 0 ⟩ = √1 ( | ↑, ↓ ⟩ + | ↓, ↑ ⟩) , (2·95)
2
| 1, −1 ⟩ = | ↓, ↓ ⟩ .
For clarity, we have suppressed the trivial quantum numbers s1,2 = 1/2 in the kets of the
coupled states, i.e., | SMS ⟩ = | ( 12 21 )SMS ⟩.
The behavior of the coupled states under transposition is interesting. As can be seen
from the explicit representation, the singlet state is antisymmetric with respect to the
interchange of the two single-particle spins. The triplet states, on the other hand, are
symmetric. This information is important when constructing antisymmetric two-particle
states in the complete two-particle Hilbert space. If the state in the spin space is symmetric
(triplet), then the state in the position and isospin space must be overall antisymmetric
to achieve an antisymmetric total state. Conversely, if the spin state is antisymmetric
(singlet), the state in the position-isospin space must be overall symmetric.
The considerations for coupling in the spin space can be directly transferred to the
isospin isospin space. For the case relevant to nucleons t = 1/2, the coupled basis | TMT ⟩ =
space
| ( 12 21 )TMT ⟩ is given by the following representation from the uncoupled basis:
| 0, 0 ⟩ = √1 ( | p, n ⟩ − | n, p ⟩) ,
2
| 1, +1 ⟩ = | p, p ⟩ ,
(2·96)
| 1, 0 ⟩ = √1 ( | p, n ⟩ + | n, p ⟩) ,
2
| 1, −1 ⟩ = | n, n ⟩ .
Regarding the symmetries of the isospin singlet and isospin triplet and the consequences
for the antisymmetry of the total state, the same applies as for the spin.
LS Coupling
As the next step, we include the spatial space and consider the full single-particle space of
the nucleon Hspatial ⊗ Hspin ⊗ Hisospin with a basis of the form | nlml ⟩ ⊗ | 21 ms ⟩ ⊗ | 21 mt ⟩.
The two-nucleon space results from the Kronecker product of two single-particle spaces
Following the previous considerations, we can couple the two single-particle orbital an-
gular momenta, the two single-particle spins, and the two single-particle isospins each to
a total orbital angular momentum L, ML , total spin S, MS , and total isospin T, MT . The
following partially coupled basis results:
To maintain clarity, the consistent use of bracket notation to illustrate the coupling scheme
is essential.
We can now go one step further and couple the total orbital angular momentum and
the total spin to a total angular momentum J, M. This leads to two-particle states in the
so-called LS coupling LS coupling
The question arises as to what symmetries the LS-coupled states exhibit with respect
to the transposition of single-particle quantum numbers. This question can be relatively
easily answered since the behavior of the CG coefficients under the interchange of the first
two columns is known (see page 30). The application of the transposition operator causes
the single-particle quantum numbers to be swapped, i.e., l1 ↔ l2 , s1 ↔ s2 , etc. In equation
(2·99), this results in the swapping of the first two columns of the CG coefficients. This can
be reversed by exploiting the symmetries of the CG coefficients, thus restoring the original
order of the angular momentum quantum numbers. Summarized:
The step from the first to the second line is simply the application of the transposition
operator. From the second to the third line, the single-particle quantum numbers (except
for the radial quantum number) are restored to their original order using the symmetries
of the CG coefficients. This results in a phase factor that provides information about the
transposition symmetry of the states.
For the special case n1 = n2 and l1 = l2 , we can read-off the symmetry of the LS-coupled
two-particle states. Antisymmetric states only arise if L+S+T is odd, since the phase factor
is then negative. As discussed in the previous section, for a state that is antisymmetric
in the spin-isospin space (S + T odd), the spatial part must be symmetric (L even). For
symmetric spin-isospin states (S + T even), the spatial part must be antisymmetric (L odd).
In the general case, i.e., n1 , n2 , the states (2·100) have to be explicitly antisymmetrized.
The spatial part of the states must be symmetrized (antisymmetrized) if the spin-isospin
part is antisymmetric (symmetric), i.e., if S + T is odd (even).
LS coupling is particularly convenient when it comes to calculating matrix elements of
two-particle operators, such as the NN interaction. We will make extensive use of this in
this context.
jj Coupling
Another scheme for the two-step coupling of single-particle angular momenta to form a
two-particle total angular momentum J, M can also be constructed. In the first step, the
orbital angular momenta and spins of the individual nucleons are coupled to form single-
particle total angular momenta j1 , m1 and j2 , m2 :
In the second step, the single-particle total angular momenta are coupled to form the
jj-coupling two-particle total angular momentum, hence the name jj-coupling,
= | n1 l1 j1 ; n2 l2 j2 ; JM; TMT ⟩ ,
where the last line defines an abbreviated notation omitting the trivial quantum numbers.
As in the case of LS coupling, the symmetry of the coupled states can be derived from the
properties of the CG coefficients. The interchange of the single-particle quantum numbers
leads to:
For the special case n1 = n2 , (J − j1 − j2 + T − 1) must be odd for the jj-coupled state to
be antisymmetric. In the general case, antisymmetric jj-coupled states can be produced
through explicit antisymmetrization.
The jj-coupling plays a major role in many-body methods that are built on s basis
expansion using Slater determinants of j-coupled single-particle states (cf. chapters 7 and
8). Consider, e.g., the NN interaction that will typically be fed into these many-body
calculations in the form of jj-coupled two-body matrix elements
⟨ n1 l1 j1 ; n2 l2 j2 ; JM; TMT | V̂NN | n′1 l′1 j′1 ; n′2 l′2 j′2 ; J′ M′ ; T′ M′T ⟩ . (2·105)
Since the two-body states carry explicit total angular momentum quantum numbers J, M
and J′ , M′ we can make use symmetry-based selection rules that restrict the possible matrix
elements. For a rotationally invariant operator (we will discuss this in the next chapter), all
matrix elements with J , J′ or M , M′ are zero and the matrix elements are independent
of the value of M. In jj-coupling we can make use of this and calculate and store only the
non-vanishing matrix elements — this a a tiny subset of all possible matrix elements, i.e.,
this symmetry-based compression is very efficient. For the many-body treatment, we usu-
m-scheme ally need uncoupled, so-called m-scheme matrix elements. When starting from jj-coupled
matrix elements, it is very easy to uncoupled them numerically on-the-fly by reverting
(2·103). In this way we can make use of the symmetry-based compression but still obtain
the needed m-scheme matrix elements efficiently. All of the many-body methods we will
discuss in chapters 7 and 8 make use of this trick.
However, the direct calculation of interaction matrix elements in jj-coupling is ex-
tremely cumbersome. Typically, matrix elements are calculated in an LS-coupled scheme
and then transformed into jj-coupling. Both coupling schemes provide a complete basis
in the two-particle space, so switching from one scheme to the other represents a unitary
transformation. This recoupling can be formally written as follows: recoupling
The matrix in curly brackets describes the so-called 9j-symbol, which can be expressed as
a sum over products of CG coefficients. We will not go into further detail at this point, but
remark that there is much more to be said about angular momentum coupling.