2223 Advanced Quantum Notes
2223 Advanced Quantum Notes
PHYS6003
Pasquale Di Bari
(e-mail:P.Di-Bari@soton.ac.uk)
2
Contents
Introduction 7
1 Vector spaces 13
1.1 Definition of vector spaces . . . . . . . . . . . . . . . . . . . . 13
1.2 Linear dependence and linear independence . . . . . . . . . . 15
1.3 Dimension and basis of a vector space . . . . . . . . . . . . . . 16
1.4 Inner product . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4.1 Orthogonal vectors and orthonormal basis . . . . . . . 18
1.4.2 Inner product in terms of the components . . . . . . . 19
1.5 Dual space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.6 Gram-Schmidt procedure . . . . . . . . . . . . . . . . . . . . . 23
1.7 Useful inequalities . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.8 Linear operators . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.9 Adjoint of an operator . . . . . . . . . . . . . . . . . . . . . . 28
1.10 Hermitian and Unitary operators . . . . . . . . . . . . . . . . 30
1.11 Matrix representation of operators . . . . . . . . . . . . . . . . 31
1.12 Outer product . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
1.13 Projection operators . . . . . . . . . . . . . . . . . . . . . . . 35
1.14 Completeness relation . . . . . . . . . . . . . . . . . . . . . . 36
1.15 Unitary transformations . . . . . . . . . . . . . . . . . . . . . 39
1.16 Active and passive transformations . . . . . . . . . . . . . . . 40
1.17 Trace and determinant of an operator . . . . . . . . . . . . . . 41
1.18 The eigenvalue problem . . . . . . . . . . . . . . . . . . . . . . 42
1.18.1 The characteristic equation . . . . . . . . . . . . . . . 43
1.19 The eigenvalue problem for Hermitian operators . . . . . . . . 45
1.19.1 Non-degenerate case . . . . . . . . . . . . . . . . . . . 45
1.19.2 Degenerate case . . . . . . . . . . . . . . . . . . . . . . 46
1.19.3 Eigenvectors and eigenvalues of unitary operators . . . 47
1.19.4 Diagonalization of Hermitian matrices . . . . . . . . . 48
3
4 CONTENTS
The name of this module could be intimidating since the word Advanced
might let you think that the level of difficulty of the treated subjects is
much higher than in the 2nd year Quantum Physics module. In fact, in
some respects, it is maybe even lower! Probably it is more correct to say
that in this module the kind of technical tools are different: In quantum
physics you mainly had to solve the Schrödinger equation applied to different
problems and in most cases the solution requires a knowledge of calculus with
boundary conditions and special functions. In our case, quantum physics is
discussed to a more formal (axiomatic) and fundamental level and, as we will
see, mainly we will have to deal with symbolic operations, mainly involving
algebraic transformations in vector spaces. In our discussion, the description
of a quantum state in terms of the wave function, for example ψ(x, t) for
a one-dimensional system, will be replaced by a state vector defined in an
infinite-dimensional complex vector space. In the Dirac notation, this will be
denoted by |ψ(t)i, an object called ket that, in general, also depends on time
(though often we will imply time dependence). It will be then important, at
the beginning, to get familiar with the notion of vector spaces.
The most familiar vector space you know is certainly the three dimen-
sional Euclidean space where a generic vector ~v (e.g., position or velocity)
can be expressed in terms of basis vectors i, j, k simply as
~v = v1 i + v2 j + v3 k , (1)
where v1 , v2 and v3 are the components of ~v in the basis {i, j, k} and are real
numbers.
Another example of vectors, though quite a different kind, is given by
complex numbers. As you know a complex number z = x + i y can be
represented in the complex plane, or the Argand plane, where the real part
x takes values on the horizontal axis and the imaginary part on the vertical
axis. Operations between complex numbers can be then expressed vectorially.
7
8 CONTENTS
In these two examples the basis is discrete since the basis vectors and
the components of a generic vector can be labelled by a discrete index, for
example we could re-express in the Euclidean case i = x1 , j = x2 , k = x3 .
Therefore, they are examples of discrete vector spaces.
In general, however one can also have a continuous basis, where basis
vectors and components are labelled by a continuous parameter.
The familiar examples we mentioned can be further generalised to vector
spaces with an arbitrary number of dimensions and with components given
by complex numbers. In quantum mechanics (QM) the state of a physical
system, to which we will refer as quantum state, is described by a ket |ψi
and this can be regarded as a state vector in an infinite-dimensional complex
space satisfying certain conditions (it needs to be complete and unitary),
mathematically what is known as Hilbert space.
The basis vectors correspond to the eigenvectors of Hermitian operators.
Hermitian operators have real eigenvalues and are associated to physical ob-
servables. A measurement can be regarded as an operation h ψ |Ô| ψ i, where
Ô is an Hermitian operator. The spectrum of Hermitian operators can be
both continuous and/or discrete, therefore, the basis vectors for quantum
states will have, in general, both continuous and discrete indexes. The de-
scription of a physical system in QM in terms of state vectors is more fun-
damental and general than a description in terms of wave functions. A wave
function would correspond to components with continuous indexes (the ar-
gument of the wave function). Of course the wave function you are typically
used, and the simplest one, is the wave function in coordinate unidimensional
space, ψ(x, t). As we will discuss, in the non-relativistic limit one can simply
derive from the quantum state the usual wave function (as we will see, it will
correspond to the operation ψ(x, t) = hx|ψ(t)i). Therefore, the formalism
you are already familiar with can be fully recovered. However, if one tries
to extend the use of a wave function to the relativistic regime, then different
fundamental issues are encountered.
The easiest way to understand this is to consider the simple case of a
particle that at time t = 0 is located inside a small box centred at the
origin of the reference frame. The solution for the wave function is in this
case well known. Suppose now to open the box and let the particle diffuse
freely. One would find that at a time t the wave function is non-vanishing
to arbitrarily large distances, even for |~r| > c t, where c is the speed of light.
This implies that there would be a non-vanishing probability to measure
the particle outside the light cone of the particle initially placed in the box.
Obviously this would violate the principle on the existence of a maximum
CONTENTS 9
speed of the causal signal corresponding to the speed of light and translating
into Lorentz invariance and shows an intrinsic obstacle of generalising the
concept of wave function to the relativistic case or, in other words, there is a
clash between a wave function formulation of QM and special relativity. This
becomes even more evident in particle physics in processes where colliding
particles are destroyed and new particles are created.
The solution of these problems is provided by Quantum Field Theories
that ultimately represent the consistent way to formulate a relativistic exten-
sion of QM to describe Nature (barring gravity!) but this goes well beyond
the scope of the module. However, a formulation of QM in terms of quan-
tum states described by kets provides the correct fundamental framework for
quantum field theories and, therefore, it represents the first necessary step
toward a relativistic formulation of QM. Therefore, this formalism should
be regarded (to our current knowledge!) as the fundamental underlying de-
scription of the state of a system. It is indeed so fundamental that any
attempt to replace it with a more fundamental one has failed so far.1 We
will see an example with the attempt of hidden variables theory that has
been experimentally disproved.
For this reason the formalism of quantum states is a necessity. Finally,
let us mention that since the components of a state vector are complex, there
is a fundamental difference with respect to Classical Mechanics.
Let us mention an important feature of QM that clearly distinguishes it
from Classical Mechanics. In QM if one has quantum states corresponding to
two different outcomes of a certain observable, say A and B, the generic quan-
tum state describing the system will be given by a superposition of the two
basis vectors (the eigenstates of the operator associated to the observable)
corresponding to the two different outcomes each with its own component α
and β, explicitly
|ψi = α|Ai + β|Bi , (2)
where α and β are complex numbers (the components of the state vector
in this case). If the state is multiplied by an overall complex number, the
quantum state does not change.2 The interesting thing is that a measurement
1
Notice that even string theory relies on such a framework and it is not a theory from
which one can for example think to derive the principles of QM. There is a clear statement
on this point made by to E. Witten during a conversation about the great synthesis that
took place in 1990 at the ICTP (Trieste) with Nobel laureate Abdus Salam [1].
2
Indeed as we will see, a quantum state is described by an infinite set of state vectors
called ray. It is like if the direction of the state vector defines the quantum state while its
length is irrelevant. We will be back on this point in more detail.
10 CONTENTS
will continue to give as a result just either A or B but the outcome cannot be
predicted for the single measurement one can only know that the probability
of obtaining A will be given by |α|2 /(|α|2 + |β|2 ) and the probability of
obtaining B will be given by |β|2 /(|α|2 + |β|2 ). This is very different from
Classical Mechanics, where in general the combination of two physical states
results in general into a new mixed state with new possible outcomes than
A and B and this is, at least in principle, predictable.
Bibliography
11
12 BIBLIOGRAPHY
Chapter 1
Vector spaces
13
14 CHAPTER 1. VECTOR SPACES
The operation of sum of vectors must satisfy the following linear properties:
• For any generic vector |ui there is an inverse vector | − u i such that
| u i + | − u i = | 0 i.
From these axioms it follows that 0 |ui = |0i and | − u i = −| u i (ps n. 1).
Note: In the following, we will denote the null vector simply by 0, i.e.,
replacing | 0 i → 0.2
In order to simplify the notation, it will prove convenient in the following
to denote the product of a vector | u i by a scalar α simply by | α u i ≡ α | u i.
It will also prove convenient to denote the sum of two generic vectors | u i
and | v i by | u + v i ≡ | u i + | v i.
The Euclidean space is a particular example of three-dimensional real
vector space. In this case given a (orthonormal) basis (reference frame) with
2
This is because we will have in store the notation | 0 i to denote the ground state
(the state with minimum energy) of systems such as the harmonic oscillator, as usually
in physics books such as those by Dirac or Sakurai (see references). The notation | 0 i
for the null vector is encountered in more formally, mathematically oriented, books such
as [2]. The ground state of the fundamental theory is called the vacuum state and its
identification and properties have very important physical consequences. Currently this is
unkown since we do not know yet the fundamental theory of Nature.
1.2. LINEAR DEPENDENCE AND LINEAR INDEPENDENCE 15
unit basis vectors i, j and k, a generic vector |vi, that in the specific case of
Euclidean space is commonly denoted by ~v , can be expressed as
|vi = ~v = v1 i + v2 j + v3 k . (1.1)
where we have expressed it in terms of unit column vectors that provide the
simple basis for column vectors.
We need to generalise to our V the definition of basis and components.
1,n
X ai
|j i = − |ii. (1.4)
i6=j
a j
16 CHAPTER 1. VECTOR SPACES
The coefficients ui are the components of | u i in the given basis and they are
unique.3
Proof. Let us prove it by contradiction, supposing that there is some
vector | v i which cannot be expressed as a linear combination of basis vectors.
In that case we would have found a set of n + 1 linearly independent vectors,
the n basis vectors plus | v i, but this would be in contradiction with the
definition of n-dimensional vector space.
Moreover, we can also prove by contradiction that the components need
to be unique in the given basis. If they were not, then one could write
n
X
|ui = u0i | i i , (1.6)
i=1
where u0i 6= ui at least for one particular value of i. However, subtracting the
two different linear combinations, more precisely subtracting Eq. (1.6) from
Eq. (1.5), one would find
n
X
(ui − u0i ) | i i = 0 , (1.7)
i=1
but this would imply ui = u0i for all i, since otherwise the basis vectors would
be linearly dependent.
Notice that even though the components of the vector | u i are unique for
a given basis, they change in a basis transformation. Explicitly, if we make
3
The components are scalars and, therefore, in a complex vector space they are complex
and in a real vector space, such as the Euclidean space, they are real.
1.4. INNER PRODUCT 17
is also independent of the basis. When the basis is specified, relations among
kets can be translated into relations among the components. For example,
the relation (1.8) would translate in the basis | i i simply in a analogous
relation among the components
wi = ui + vi , (1.9)
implying that the components of the sum of two vectors are given by the
sums of their components. In a new basis {| i0 i}, the components would
change in a way that Eq. (1.31) would simply transform into
for some choice of the coefficients ai ’s. Taking the inner product of each term
with a vector | j i, where j is a particular value of i, and considering that
from axiom (iii) one has h j | 0 i = 0, one has
Xn
ai h j | i i = 0 . (1.14)
i=1
Since the vectors are mutually orthogonal we have for any i, except for i = j,
that h j | i i = 0, obtaining aj = 0. Since the procedure can be iterated for
any j, one arrives to the conclusion that all ai = 0 and, therefore, that the
set {| i i} is necessarily linear independent.
However, notice that the opposite is not necessarily true: vectors in a
linearly independent set are not necessarily mutually orthogonal.6
Definition. A set of basis vectors all with unit norm, which are pairwise
orthogonal is called an orthonormal basis.
This implies that if the vectors | 1 i, | 2 i, . . . , | n i form an orthonormal
basis, then one has (for i, j = 1, . . . , n)7
h i | j i = δij . (1.15)
6
It is easy to get convinced on this point thinking of vectors in the Euclidean space:
one can clearly have three linearly independent vectors that are not mutually orthogonal.
7
The symbol δij is the Kronecker delta defined by δij = 1 for i = j and δij = 0 for
i 6= j.
8
Notice that we replaced the dummy index i in Eq. (1.18) with j, since it is better not
to use the same dummy index in the same formula, it can easily lead to incorrect results.
20 CHAPTER 1. VECTOR SPACES
Using the linearity of the inner product in the second vector (axiom (iii))
and the anti-linearity in the first vector (see Eq. (1.11)) one arrives to the
expression
X n X n
hu|vi = u?i vj h i | j i . (1.19)
i=1 j=1
Notice that the inner products between basis vectors, the h i | j i’s , depend
on the basis. Specifically, for an orthonormal basis one can use Eq. (1.15) to
express them. In this way, under the action of δij , the double sum collapses
into a single sum. Therefore, for an orthonormal basis, one simply obtains:
n
X
hu|vi = u?i vi . (1.20)
i=1
an expression consistent with the axiom (ii) implying that the norm of a
vector is a nonnegative real number.
In a given basis a vector is specified by its components and, therefore,
one can represent a ket | u i in terms of a column vector formed by its com-
ponents:9
u1
u2
|ui → u ≡ ... .
(1.22)
un
The inner product h u | v i of two kets | u i and | v i can then be obtained as the
matrix product of the transpose conjugate of the column vector representing
| u i with the column vector representing | v i, explicitly:
v1
n
X v2
hu|vi = u?i vi = u?1 u?2 . . . u?n ... .
(1.23)
i=1
vn
9
In the following the (short) arrow symbol → will always stand for ‘is represented in
the given basis by’.
1.5. DUAL SPACE 21
Notice that denoting the conjugate transpose or adjoint of u by u† ≡ (u?1 u?2 . . . u?n ),
we can also express, more compactly, the inner product in terms of a matrix
product:
X n
hu|vi = u?i vi = u† v . (1.24)
i=1
• The set of all bras form the so-called dual space of V denoted by V ? ;
• Any basis of kets {| i i} has its adjoint basis of bras {h i |} and vice-versa;
10
One can now finally appreciate the origin of the name ‘ket’: taking the inner product
with a bra corresponds to form a bra(c)ket!
11
It is the analogous of forming a contravariant vector from a covariant vector in rela-
tivity. However, notice that in relativity the squared length Aµ Aµ of a generic 4-vector
Aµ can also be negative!
22 CHAPTER 1. VECTOR SPACES
• If one has a ket | αu i, where α is a scalar, then its adjoint bra is given
by12
h αu | = h u | α? , (1.25)
since if one has a column with components {α ui } then its adjoint will
be a row with components {α? u?i }, explicitly:
α u1
α u2 adjoint
| αu i → ←→ α? u?1 α? u?2 . . . α? u?n ← h αu | .
...
α un
(1.26)
• From any relation among kets, such as
α|ui = β |vi + γ |wi, (1.27)
one can obtain the adjoint expression replacing each term (a scalar
multiplied by a ket) with its adjoint (a bra multiplied the complex
conjugate of the scalar) and vice-versa. Explicitly, in the previous
example, one obtains:
h u | α? = h v | β ? + h w | γ ? . (1.28)
This can be expressed saying that to take the adjoint of an equation
relating kets (bras), one needs to replace every ket (bra) with the cor-
responding bra (ket) taking the complex conjugate of all coefficients.
• If the expansion in components of a ket | u i in the given basis is given
by
X n
|ui = ui | i i , (1.29)
i=1
then taking the adjoint of this relation gives us the expansion of a bra
in the adjoint basis:
Xn
hu| = h i | u?i . (1.30)
i=1
12 ? ?
Notice that one has h u | α = α h u |. Usually there is some convenience in writing
the scalar on the right of a bra analogously to what must be done with operators since
as we will see the multiplication by a scalar α can be regarded as applying the operator
ˆ where Iˆ is the identity operator, to the ket. However, as we will see, there are also
αI,
situations where it is more convenient to write the scalar on the right of the ket and on
the left of the bra.
1.6. GRAM-SCHMIDT PROCEDURE 23
• By taking the inner product of the expansion (1.29) with a basis bra
h j |. Using the orthonormality of the basis implying h j | i i = δij , one
immediately finds that the components can also be written as
uj = h j | u i , (1.31)
so that one can also write the expansion of a ket | u i as
n
X
|ui = hi|ui|ii. (1.32)
i=1
Taking the adjoint of this expression one immediately finds the corre-
sponding expression for the bra
n
X
hu| = hi|hu|ii, (1.33)
i=1
0 | 20 i
| 2 i −→ | 2 i = 0 . (1.37)
||2 ||
| e3 i −→ | 30 i = | e3 i − | 1 ih 1 | e3 i − | 2 ih 2 | e3 i; (1.38)
n) Finally, with the last n-th step one transforms | en i → | n i using the
mutually orthogonal kets with unit norm | 1 i, . . . , | n − 1 i obtained in
previous steps. It can again be decomposed into two sub-steps:
Notice that since the initial basis {| ei i} is, by definition of basis, a linearly
independent set, one has | i0 i =
6 0 for all i = 1, . . . , n. This is because if
by subtracting all parallel components to | 1 i, . . . , | m − 1 i one obtains 0 for
1.7. USEFUL INEQUALITIES 25
2 Re(h u | v i) ≤ h u | u i + h v | v i . (1.43)
Proof. It is quite easy to prove it. From axiom (ii) we can impose
hu − v|u − vi ≥ 0, (1.44)
and then decomposing the left-hand side using properties of the inner prod-
uct, one obtains
h u | u i + h v | v i − h v | u i − h u | v i ≥ 0. (1.45)
hv|ui
hw| = hv| − hu|, (1.48)
hu|ui
one finds
h u | v i? h u | v i h u | v i h v | u i h v | u i h u | v i
h v | v i− − + h u | u i ≥ 0 , (1.49)
||u||2 ||u||2 ||u||4
and from this, using h u | u i = ||u||2 so that last two terms cancel out and
considering that h u | v i? h u | v i = |h u | v i|2 , with easy algebraic steps one
immediately finds the Cauchy-Schwarz inequality Eq. (1.46).
Theorem: Triangle inequalities. The first states that the norm of the sum
of two vectors cannot exceed the sum of their norms, explicitly:
The second states that the norm of the sum of two vectors cannot be lower
than the absolute value of the difference of the norms, explicitly:
The proofs are quite simple, they are left as problems in problem-sheet n.2.
Ô
h v | −→ h ve0 | = h v |Ô , (1.53)
where notice that h ve0 | does not coincide in general with the adjoint of | v 0 i
and that is why we did not denote it by h v 0 |. We will see in the next section
how to obtain the bra adjoint of | v 0 i = | Ov i.
1.8. LINEAR OPERATORS 27
We will only deal with linear operators, defined by the following rules:
Ôα| u i = α Ô| u i , (1.54)
Ô (α| u i + β| v i ) = αÔ| u i + β Ô| v i , (1.55)
h u |α Ô = h u |Ô α , (1.56)
(h u |α + h v |β) Ô = h u |Ôα + h v |Ôβ . (1.57)
Definition. The simplest operator is the identity operator, denoted by I, ˆ
transforming any vector into itself, so that explicitly for every ket | u i one
has
ˆ ui = |ui,
I| (1.58)
and for every bra h u |
h u |Iˆ = h u | . (1.59)
Linear operators have the advantage that once their action on the basis
vectors is known, then one can immediately determine their action on any
vector. This is simple to show. Suppose one has for each basis vector | i i in
a basis {| i i} ≡ {| 1 i, | 2 i . . . , | n i}
Ô| i i = | i0 i , (1.60)
then applying the operator to the expansion of a generic ket | u i in the basis
{| i i} (see Eq. (1.29)) and using the linearity of Ô, one immediately finds
n
X
Ô| u i = Ôui | i i , (1.61)
i=1
n
X
= ui Ô | i i , (1.62)
i=1
n
X
= ui | i0 i . (1.63)
i=1
The product Ô2 Ô1 of two operators Ô1 and Ô2 is also an operator. Its action a
generic vector | u i is simply given by applying sequentially the two operators,
in a way that one has:
Ô2 Ô1 | u i = Ô2 (Ô1 | u i) = Ô2 | O1 u i = Ô2 | u0 i = | u00 i , (1.64)
that can be also more explicitly expressed as the double transformation:
Ô Ô
1
| u i −→ | u0 i = Ô1 | u i ≡ | O1 u i −→
2
| u00 i = Ô2 | u0 i = Ô2 Ô1 | u i ≡ | O2 O1 u i .
(1.65)
28 CHAPTER 1. VECTOR SPACES
An important point is that in general the order of the two operators matters,
in a way that defining the commutator of the operator Ô with P̂ as
h i
Ô, P̂ ≡ Ô P̂ − P̂ Ô , (1.66)
If the commutator vanishes, then it is said that the two operators commute
with each other.
If one consider three operators Ô, P̂ and Q̂, then one has these two useful
identities:
h i h i h i
Ô, P̂ Q̂ = P̂ Ô, Q̂ + Ô, P̂ Q̂ , (1.68)
h i h i h i
Ô P̂ , Q̂ = Ô P̂ , Q̂ + Ô, Q̂ P̂ . (1.69)
Definition. Given an operator Ô, its inverse Ô−1 is defined as that oper-
ator such that
Ô−1 Ô = Ô Ô−1 = Iˆ . (1.70)
Notice that not every operator has an inverse. If two operators Ô and P̂
both possess an inverse operator, then the inverse of their product is given
by the product of their two inverses in reverse order, explicitly:
This can be easily verified: (Ô P̂ )−1 Ô P̂ = P̂ −1 Ô−1 Ô P̂ = Iˆ and likewise one
ˆ
finds Ô P̂ (Ô P̂ )−1 = I.
h u |Ô† = h Ou | , (1.72)
1.9. ADJOINT OF AN OPERATOR 29
Here again let us notice that the scalars αi? might well be placed also to the
left of the bras but it should now be clear that placing them to the right
of bras highlights their nature of special operators, resembling the way how
adjoint operators are written.
We can summarise the rules that bring from one relation to its adjoint
saying that one has to operate the following substitutions also holding in
reverse order:
(i) Ô ←→ Ô† ;
(ii) | . . . i ←→ h . . . | ;
(iii) αi ←→ αi? .
Û Û † = Iˆ . (1.80)
Since Û and Û † are inverse of each other, then, necessarily from Eq. (1.70),
ˆ A unitary operator is the analogous of a complex
one also has Û † Û = I.
number with unit modulus, e.g., of the kind z = eiϕ , satisfying z z ? = 1.
A very important property of unitary operators is the following.
Theorem. Unitary operators preserve the inner products of vectors.
Proof. Consider a unitary operator Û and two generic kets | u i and | v i
that, under the action of Û , get respectively transformed into
| u0 i = Û | u i and | v 0 i = Û | v i . (1.81)
h u0 | v 0 i = h U u | U v i = h u |Û † Û | v i = h u | v i . (1.82)
and of course, like the components of a vector, the matrix elements are also
basis-dependent.14 As we will show explicitly in a moment, the j-th column
of O is formed by the components of the ket | j 0 i = Ô| j i, the transformed
basis vector | j i, in the given basis.
Let us see the importance of knowing the matrix elements in concrete
calculations. We have seen that once we know the set of transformation rules
of an operator on basis vectors, then one can derive easily the transformation
rule for a generic vector and we obtained the result
n
X n
X
0
| u i = Ô| u i = uj Ô | j i = uj | j 0 i . (1.85)
j=1 j=1
14
On the other hand, it should be clear that abstract operators such as Ô, like abstract
vectors, are not basis-dependent.
32 CHAPTER 1. VECTOR SPACES
Notice that in the sum we deliberately wrote Oij uj instead of uj Oij , since
one can immediately recognise that Eq. (1.87) is equivalent to the matrix
form relation:
u01 O11 O12 . . . O1n u1
u0 O21 O22 . . . O2n u2
2
.. = .. .. .. , (1.88)
.. . .
. . . . . .
u0n On1 On2 . . . Onn un
or, more compactly, u0 = O u. Notice that this relation is nothing else than
the representation, in the given basis, of the abstract relation | u0 i = Ô| u i
defining the operator Ô.
Example. If one considers a basis vector | j i, in this case one obtains as
a transformed basis vector
O11 . . . O1j . . . O1n 0 O1j
. .
O21 . . O2j . . . O2n .. O.2j
. .. .. 1 = .. ,
.. . . (1.89)
. . .
.. ... .. .. ..
.
On1 . . . Onj . . . Onn 0 Onj
showing that indeed the elements of matrix in a given j-th column corre-
sponds to the components of the transformed basis ket | j i.
1.11. MATRIX REPRESENTATION OF OPERATORS 33
This result was somehow expected: the matrix representing the adjoint
of Ô is the adjoint matrix (transpose complex conjugated) of the matrix
representing Ô.
| u ih v | , (1.100)
| w0 i = | u ih v | w i . (1.101)
| v ih v |
P̂v ≡ (1.102)
hv|vi
P̂i = | i ih i | . (1.105)
The relation (1.103) for the transformed vector | w0 i under the action of
P̂i on a generic vector | w i specialises then into
P̂
i
| w i −→ | w0 i = P̂i | w i = | i ih i | w i = wi | i i , (1.106)
P̂
v
h w | −→ h w0 | ≡ h w |P̂i = h w | i i h i | = wi? h i | . (1.107)
Taking the product of two projection operators onto different basis kets
| i i and | j i, one has:
This implies that P̂i P̂j = 0 if i 6= j, while for i = j one obtains P̂i2 = P̂i .
This result was expected since, if one projects for a second time on the same
vector, then necessarily the result is the same since the vector was already
projected. On the other hand if one projects the second time on an orthogonal
vector, then the parallel component to this orthogonal vector has necessarily
to vanish and, therefore, one obtains the zero vector.
If we calculate the matrix elements of P̂i , we obtain
where the representing matrix has just one nonzero element along the diag-
onal in the (i, i) entry.
We have seen that the components ui can be written in terms of inner prod-
ucts as ui = h i | u i. Therefore, the expansion (1.113) can be also written
as !
X n n
X n
X
|ui = | i ih i | u i = P̂i | u i = P̂i | u i . (1.114)
i=1 i=1 i=1
This expressions is telling that the operator in the brackets is leaving un-
changed the ket | u i but since the ket | u i is generic, then necessarily the
only operator that leaves unchanged any vector is the identity and therefore
we proved the theorem.
The completeness relations can be used to obtain different useful results
since it can be inserted inside an abstract relation to get its matrix form. We
give here below two important applications.
Theorem. Given an orthonormal basis {| i i}, where i = 1, . . . , n, any
generic operator Ô can be decomposed in terms of outer products | i ih j |
as15
1,n
X
Ô = Oij | i ih j | . (1.115)
i,j
Proof. It can be shown either simply verifying that taking the matrix ele-
ments of both sides one indeed obtains Oij = Oij or, alternatively, one can
use twice the completeness relation, in the following way:
Ô = Iˆ Ô Iˆ (1.116)
n
! n
!
X X
= | i ih i | Ô | j ih j | (1.117)
i=1 j=1
1,n
X
= | i ih i |Ô| j ih j | (1.118)
i,j
1,n
X
= Oij | i ih j | . (1.119)
i,j
15
P1,n Pn Pn
We denoted the double sum by the compact notation: i,j ≡ i=1 j=1 .
38 CHAPTER 1. VECTOR SPACES
where the unity in the (i, j) entry is the only nonzero entry. Therefore, the
relation (1.116) in matrix form reads as
0 ... ... 0
O11 O12 . . . . . . . . . O1n
O21 O22 . . . . . . . . . O2n .. .. ..
. . .
.. .. .. ..
. . . . X 1,n
.
.
. ... 1 ...
.. = Oij . .
.
.. .. ..
.
. . . .. ..
i,j
. . . .
..
..
.. .. .. ..
. .
On1 On2 . . . . . . . . . Onn 0 ... ... 0
(1.121)
By using the completeness relation, we can also find an expression for the
matrix elements of a product of two generic operators Ô and P̂ in terms of
products of matrix elements of the two operators:
(O P )ij = h i |Ô P̂ | j i (1.122)
n
!
X
= h i |Ô | k ih k | P̂ | j i (1.123)
k=1
n
X
= h i |Ô| k ih k |P̂ | j i (1.124)
k=1
Xn
= Oik Pkj . (1.125)
k=1
As one could expect the result is that the matrix elements of the product of
two operators are given by the sum of the product of the matrix elements of
1.15. UNITARY TRANSFORMATIONS 39
n
X
| i0 i = | k ih k | i0 i (1.126)
k=1
n
X
= | k ih k |Û | i i (1.127)
k=1
n
X
= | k iUki . (1.128)
k=1
n
X
0
hj | = h j 0 | ` ih ` | (1.129)
`=1
n
X
= h j |U † | ` ih ` | (1.130)
`=1
Xn
†
= Uj` h`|. (1.131)
`=1
40 CHAPTER 1. VECTOR SPACES
Since we are assuming that the new basis is orthonormal, one can impose
δji = h j 0 | i0 i (1.132)
1,n
X †
= Uj` Uki h ` | k i (1.133)
k,`
n
X †
= Ujk Uki , (1.134)
k=1
where in the second step we replaced the two expansions written before. The
last relation is nothing else than the matrix form of the abstract relation
ˆ so that we can conclude that Û is a unitary operator.
Û † Û = I,
Therefore, we conclude that the operator Û that induces the transforma-
tion of the basis vectors from one orthonormal basis to another, is necessarily
unitary. Conversely notice that the same proof also shows that given a uni-
tary operator, this can be used to generate a new orthonormal basis starting
from an initial one.
Û
h u |Ô| v i −→ h Û u |Û ÔÛ † | Û v i = h u |Û † Û ÔÛ † Û | v i (1.139)
= h u |Ô| v i .
This definition is perfectly consistent with the fact that the operator is an
object in abstract and, therefore, basis-independent. Using the completeness
relation, one can indeed easily
Pprove that Tr(Ô)Pis a basis invariant quantity
(PS n.4), so that Tr(O0 ) = ni=1 h i0 |Ô| i0 i = ni=1 h i |Ô| i i = Tr(O). One
16
In general if the active transformation is induced by Û and the passive transformation
by some second generic unitary operator Ŵ , one would have
Û ,Ŵ
h u |Ô| v i −→ h Û u |Ŵ † ÔŴ | Û v i = h u |Û † Ŵ † ÔŴ Û | v i . (1.138)
can also prove that Tr(ÔP̂ ) = Tr(P̂ Ô). From this second result one can
also prove that, give a unitary operator Û and a generic operator Ô, one
has Tr(Û † Ô Û ) = Tr(Ô). This provides another way to show that Tr(Ô) is
basis invariant, since we have seen that a basis transformation is realised by
a unitary transformation.
Definition. The determinant of an operator Ô, denoted by det(Ô), is
given by the determinant of the matrix representing the operator in any
given basis basis {| i i}. As in the case of the trace of an operator, also
the determinant of an operator is basis invariant and, therefore, it can be
calculated in any basis. The proof is analogous to the proof that the trace of
an operator is invariant since one has det(O0 ) = det(U † OU ) = det(O), where
O0 is the matrix representation of Ô in a different basis {| i0 i}.
Û
| u i −→ | u0 i = Ô| u i , (1.141)
in general | u0 i =
6 ω| u i, where ω can be any scalar. However, for some special
| u i and ω, it can happen that one has
Ô| u i = ω| u i . (1.142)
In this case, barring the trivial case | u i = 0, the ket | u i is called an eigenket
of Ô with eigenvalue ω. The set of eigenvalues of an operator Ô is also referred
to as its spectrum and its determination as the eigenvalue problem for the
operator Ô. Notice that, given an eigenket | u i with eigenvalue w, any other
ket α| u i, where α is a scalar, is also an eigenket with the same eigenvalue.
Example. Consider the identity operator I. ˆ In this case one has for any
generic ket | u i
ˆ ui = 1|ui,
I| (1.143)
implying that any ket is an eigenket of Iˆ with eigenvalue ω = 1.
Example. Consider the projection operator P̂v = | v ih v | on a generic
normalised ket | v i.
• Any ket | αv i is an eigenket of P̂v with eigenvalue ω = 1:
P̂v | αv i = | v ih v | αv i = | αv ih v | v i = | αv i . (1.144)
1.18. THE EIGENVALUE PROBLEM 43
• Any ket that does not fall in one of the two previous classes, i.e., any
linear combination α| v i + β| v⊥ i with both α and β 6= 0, is not an
eigenket of P̂v .
Since these three classes include all kets in the vector space, we found all
eigenkets and eigenvalues of P̂v .
In general, the identification of the eigenvalues and eigenkets of a generic
operator does not proceed so simply and one has to follow a more involved
procedure that we are going to discuss.
In this way we obtained the representation of Eq. (1.142) in the given basis.
In matrix form the Eq. (1.148) can then be written as
O11 . . . . . . O1n u1 u1
.. ..
.
.. u u2
. . 2
.. = ω .. ⇔ O u = ω u . (1.149)
. . .
.. .. .. . .
On1 . . . . . . Onn un un
17
In practice this means taking the inner product side by side with some basis vector.
44 CHAPTER 1. VECTOR SPACES
(O − ω I) u = 0 . (1.150)
u = (O − ω I)−1 0 = 0 , (1.151)
that is absurd since we are requiring that u is not the null vector. This implies
that the solutions of Eq. (1.150) have necessarily to be found imposing that
the matrix (O − ω I) is non-invertible and this, from theory matrix, implies
that has to be singular, i.e., it must have a vanishing determinant, explicitly:
det(O − ω I) = 0 . (1.152)
h ω |Ô| ω i = ω h ω | ω i (1.156)
h ω |Ô† | ω i = ω ? h ω | ω i . (1.157)
Since we are assuming Ô† = Ô, then subtracting the last two equations side
by side leads immediately to ω = ω ? .
Another important property of Hermitian operators is the following the-
orem.
Theorem. The normalised eigenvectors of a Hermitian operator are mutu-
ally orthogonal and form an orthonormal basis, called eigenbasis. This basis
is unique if the operator is non-degenerate, otherwise there is an infinite
number of choices.
Ô| ωi i = ωi | ωi i (1.158)
46 CHAPTER 1. VECTOR SPACES
and
Ô| ωj i = ωj | ωj i . (1.159)
We can now take the inner product with h ωj | and h ωi | respectively, obtaining
h ωj |Ô| ωi i = ωi h ωj | ωi i (1.160)
and
h ωi |Ô| ωj i = ωj h ωi | ωj i . (1.161)
We can now take the adjoint of the latter, and using again Ô† = Ô we obtain
h ωj |Ô| ωi i = ωj h ωj | ωi i , (1.162)
where we used, from the previous theorem, that ωj? = ωj . We can now
subtract from this last equation the expression (1.160) side by side obtaining
0 = (ωj − ωi ) h ωj | ωi i . (1.163)
Ô| ω1 i = ω | ω1 i (1.164)
Ô| ω2 i = ω | ω2 i . (1.165)
Any linear combination of the two eigenkets is clearly still an eigenket since
Ô (α | ω1 i + β | ω2 i) = α ω | ω1 i + β ω | ω2 i
= ω (α | ω1 i + β | ω2 i) , (1.166)
20
For a general proof, see [2].
1.19. THE EIGENVALUE PROBLEM FOR HERMITIAN OPERATORS47
for any α and β. Varying continuously α and β, the linear combination will
fill the whole two-dimensional subspace spanned by | ω1 i and | ω2 i, whose
elements are all eigenkets with eigenvalue ω. This subspace is referred to as
an eigenspace of Ô with eigenvalue ω. Obviously there exists an infinity of
choices of orthonormal pairs | ω10 i and | ω20 i that can be obtained from | ω1 i
and | ω2 i by a rigid rotation of the kind
[N̂ , N̂ † ] = 0 . (1.168)
We have seen that any operator can be expanded in a generic basis in the
form (1.115). In the eigenbasis this expansion becomes particularly simple:
n
X
Ô = ωi | ωi ih ωi | . (1.172)
i=1
that in the new basis the matrix elements are given by the matrix elements
of the operator Û † ÔÛ . Therefore, this result is equivalent to say that for
any Hermitian operator Ô, represented in some orthonormal basis by a ma-
trix O, there always exists a unitary transformation Û represented by a
matrix U such that (U † O U )ij = wi δij . This is equivalent to say that the
matrix representing the operator Û † ÔÛ is diagonal. In terms of a passive
transformation one can say that there exists a unitary operator Û diagonal-
ising Ô, i.e., that transforms Ô to Ô0 = Û † Ô Û in a way that the matrix
O0 = U † O U is diagonal. In both cases, either Û is interpreted as an active
or a passive transformation, this implies that any Hermitian matrix O can
be diagonalised by some unitary matrix U and the problem of finding the
matrix that diagonalises O is equivalent to solving its eigenvalue problem
since, from Eq. (1.174), we can see that the matrix elements of U are given
by
Uij = h i | ωj i , (1.175)
Ô| λi i = ωi | λi i . (1.179)
• Basis vectors. In the discrete case the basis vectors are labelled by
a discrete index and are denoted by | i i. In the continuous case the
discrete index i is replaced by a continuous real variable x ∈ [a, b] and
basis vectors are denoted by | x i:
| i i ←→ | x i . (1.180)
Notice that, if in the discrete case one can represent a ket with a column
vector with n components, now one can think of the wave function as
a column vector with infinite elements (do not try to write it down!).
• Inner product. Similarly to the vector expansion, now the sum on
products of the components of the two vectors is replaced by an integral
over x:
X n Z b
?
hv|ui = vi ui ←→ h ϕ | ψ i = dx ϕ? (x) ψ(x) . (1.182)
i=1 a
21
The variable x is not necessarily denoting position.
52 CHAPTER 1. VECTOR SPACES
and from this one immediately derives the resolution of the identity
Eq. (1.188).
1.20.2 Operators
We have now to see how operators change in the continuous case. As in the
discrete case, they transform a vector into another vector,
Ô
| ψ i −→ | ψ 0 i ≡ | Ôψ i = Ô| ψ i . (1.190)
Given a basis {| x i}, one can again calculate the matrix elements of the
operator Ô:
Oxx0 = h x |Ô| x0 i , (1.191)
1.20. GENERALIZATION TO INFINITE DIMENSIONS 53
and one can still think of a representation matrix where the entries are given
by the matrix elements and that now will have an infinite number of rows
and columns.
The simplest operator one can introduce in the basis {| x i} is the operator
x̂ defined as that Hermitian operator with eigenkets given by the same basis
vectors | x i and with x as eigenvalue,
x̂ | x i = x | x i . (1.192)
Using the resolution of identity Eq. (1.188), we find (omitting the integration
limits in the integral)
ˆ ψi
h x |x̂| ψ i = h x |x̂ I| (1.195)
Z
0 0 0
= h x | x̂ dx | x ih x | | ψ i (1.196)
Z
= dx0 h x | x̂ | x0 ih x0 | ψ i (1.197)
Z
= dx0 x0 δ(x − x0 ) h x0 | ψ i (1.198)
= x ψ(x) . (1.199)
The result is then simply that the wave function is multiplied by x. This
result is expressed saying that in the x̂-basis the operator x̂ is represented by
x, i.e., x̂ → x.
Definition. Another important example of operator is the differential
operator D̂ defined as that operator such that22
dψ
h x |D̂| ψ i ≡ , (1.200)
dx
22
This can be also expressed saying that in the x̂-basis D̂ is represented by d/dx or
D̂ → d/dx.
54 CHAPTER 1. VECTOR SPACES
Comparing now this expression with the definition of delta function that, as
we noticed, samples a function in a specific point,
Z b
dx0 δ(x − x0 ) ψ(x0 ) = ψ(x) , (1.203)
a
we find
dψ
h x |D̂| x0 iψ(x0 ) = δ(x − x0 ) , (1.204)
dx0
and since ψ is generic, one finds for the matrix elements of D̂:
d
h x |D̂| x0 i = δ(x − x0 ) . (1.205)
dx0
Theorem. The operator D̂ is not Hermitian but the operator
k̂ ≡ −i D̂ (1.206)
is Hermitian.
Proof. In order to prove it we need to show that
Using the definition of inner product in the infinite-dimensional case Eq. (1.182),
this is equivalent to:
Z b Z b ?
? dψ(x) ? dϕ(x)
dx ϕ (x) −i = dx ψ (x) −i . (1.208)
a dx a dx
1.20. GENERALIZATION TO INFINITE DIMENSIONS 55
Integrating the left-hand side by parts and conjugating the integral in the
right-hand side, one then obtains
b b
dϕ? (x) dϕ? (x)
Z Z
−i [ϕ ?
(x) ψ(x)]ba +i dx ψ(x) =i dx ψ(x) , (1.209)
a dx a dx
and from this one finally finds that k̂ is Hermitian if the surface term vanishes,
i.e., if
[ϕ? (x) ψ(x)]ba = 0 . (1.210)
This is interesting, since it shows that in the infinite-dimensional case the
condition for k̂ to be Hermitian is stronger than in the discrete case, since it
also involves the behaviour of the wave functions at the end points a and b.
In QM we will be interested to the case a = −∞ and b = +∞ and in this
case the condition (1.210) is known as square integrability. This is a condition
that wave functions need to satisfy to be physical, i.e., to describe physical
quantum states. In the following, we will then assume square integrability
so that k̂ can be considered an Hermitian operator.
Let us now solve the eigenvalue problem for k̂. We need to find the
eigenvectors | k i satisfying
k̂| k i = k | k i . (1.211)
This can be done in the basis {| x i}. Following the usual procedure, we can
project Eq. (1.211) on the x̂ eigenvectors finding
h x |k̂| k i = k h x | k i (1.212)
and, using once more the resolution of identity in the left-hand side, one finds
Z +∞
dx0 h x |k̂| x0 i ψk (x0 ) = k ψk (x) , (1.213)
−∞
where the eigenvalue k can be any real number. The free parameter A can
be fixed imposing a normalization condition. To this extent, let us calculate
the inner products
Z +∞ Z +∞
0 0 0
hk|k i = dx h k | x ih x | k i = A 2
dx e−i(k−k ) x = 2π A2 δ(k − k 0 ) ,
−∞ −∞
(1.217)
where we used a well known representation√of the delta function. The most
common and sensible choice is then A = 1/ 2π so that finally we have that
the eigenkets of k̂ in the basis of x̂ eigenkets are represented by
1
| k i → ψk (x) = √ ei k x . (1.218)
2π
where ψ(k)
e is the wave function associated to the vector | ψ i in the k̂ basis.
It is now interesting to see how the two wave functions, ψ(k)
e and ψ(x), are
related to each other. An expression of ψ(k) in terms of ψ(x) is easily derived:
e
Z +∞
ψ(k)
e ≡ hk|ψi = dx h k | x ih x | ψ i (1.220)
−∞
Z +∞
1
= √ dx e−i k x ψ(x) . (1.221)
2π −∞
1.20. GENERALIZATION TO INFINITE DIMENSIONS 57
Of course one can also take the inverse path, obtaining ψ(x) from ψ(k):
e
Z +∞
ψ(x) = h x | ψ i = dk h x | k ih k | ψ i (1.222)
−∞
Z +∞
1
= √ dk ei k x ψ(k)
e . (1.223)
2π −∞
What we have found is that the familiar Fourier transform provides the way
to switch from the x̂ basis to the k̂ basis and vice versa with the inverse
Fourier transform. Both bases can be used to expand state vectorsstate
vector in the Hilbert space.23
We can also easily calculate the matrix elements of k̂ in the same k̂ basis:
h k |k̂| k 0 i = k 0 h k | k 0 i = k 0 δ(k − k 0 ) . (1.224)
As expected we obtained exactly the same result we obtained for h x |x̂| x0 i
(c.f. Eq. (1.193)) where simply x is replaced by k. In the same way we can
see that if we calculate the matrix elements of x̂ in the k̂ basis, we obtain
the reciprocal result obtained for the matrix elements of k̂ in the x̂ basis in
Eq. (1.214). We can this time first calculate the action of x̂ on the wave
function in the k̂ basis:
Z
h k |x̂| ψ i = dx h k | x i h x |x̂| ψ i (1.225)
Z
1
= √ dx e−i k x x h x | ψ i (1.226)
2π Z
i d
= √ dx e−i k x h x | ψ i (1.227)
2π Zdk
d
= i dxh k | x i h x | ψ i (1.228)
dk
d
= i hk|ψi (1.229)
dk
d e
= i ψ(k) . (1.230)
dk
From this result one can then also find easily, proceeding analogously to what
we have done in (1.202), the elements of matrix of x̂ in the k̂ basis:
0 0 d
h k |x̂| k i = δ(k − k ) i 0 . (1.231)
dk
23
We imply ‘physical’ when we refer to the Hilbert space from now on.
58 CHAPTER 1. VECTOR SPACES
In summary we found that in the x̂ basis the operator x̂ acts simply like
x and k̂ as −id/dx on the wave function ψ(x). On the other hand, in the k̂
basis, k̂ acts as k and x̂ like id/dk: operators like x̂ and k̂, for which such
interrelationship holds, are said to be conjugate to each other.
Another interesting result is that, if one calculates the commutator, one
finds24 h i
x̂, k̂ = i . (1.232)
Notice that here x̂ and k̂ are two generic operators. In QM we will identify
x̂ with the position operator and p̂ ≡ ~ k̂ with the momentum operator. It is
now time to show how all the formalism of vector spaces we discussed in this
chapter can be applied to QM, obtaining a much more elegant and powerful
formulation than the one based on wave functions.
In this way the direct products | i i1 ⊗ | j i2 form a basis for V1⊗2 called
product basis and the dimensionality of V1⊗2 is, therefore, given by n1 n2 , i.e.,
the product of the dimensionalities of V1 and V2 .
Notice that the direct product | ψ i1 ⊗ | ψ i2 of two generic vectors belongs
to V1⊗2 but V1⊗2 also contains linear combinations of direct products that
cannot be expressed in the form | ψ i1 ⊗ | ψ i2 . Indeed if we expand | ψ i1 and
| ψ i2 in their respective bases {| i i1 } and {| i i2 },
X (1) X (2)
| ψ i1 = ai | i i1 and | ψ i2 = aj | j i2 , (1.235)
i j
If one compares with the expression (1.234) one can see that in this this case
(1) (2)
one simply has aij = ai aj , i.e., the components of the direct product can
be expressed as the product of the components of the two distinct vectors.
and the vector is said to be separable. However, this is not possible for all
states | ψ i12 and some of them simply cannot be expressed in that way: these
states are called entangled states.
We will see that in QM, a tensor product space describes a system ob-
tained combining together different subsystems, like for example two particle
states. We will see that the implications are quite astonishing.
60 CHAPTER 1. VECTOR SPACES
Bibliography
61
62 BIBLIOGRAPHY
Chapter 2
We can now finally show in this chapter how vector spaces provide the correct
mathematical formalism to formulate QM.
63
64 CHAPTER 2. THE PRINCIPLES OF QUANTUM MECHANICS
2.2 Remarks
• Suppose the system is in a quantum state described by | ψ i and that
a measurement of the observable associated to the Hermitian operator
Ω̂ is made. One can then determine the probability to find ωi as a
result of the measurement and that, correspondingly, the state vector
collapses into the eigenket | ωi i in the following way:
i) Solve the eigenvalue problem for Ω̂, finding its eigenvalues ωi and
eigenvectors | ωi i;
ii) Expand | ψ i in the eigenbasis of Ω̂:
X
|ψi = | ωi ih ωi | ψ i ; (2.2)
i
2
Again notice that any state vector obtained by | ψ i multiplying it by some scalar, de-
scribes the same quantum state since this is determined by the ray and not by a particular
state vector of the same ray.
3
It should be appreciated how this expression can be written only for Hermitian oper-
ators, since their eigenkets, as we proved, form a basis.
2.2. REMARKS 65
(iii) The probability will be then given by Eq. (2.1). Notice that if one
has a normalised vector, then h ψ | ψ i = 1 and simply
• The first principle states that the quantum state is described by a ray
α| ψ i in Hilbert space, meaning that given a state vector | ψ i describing
the quantum state, any other state vector obtained by | ψ i multiplying
it by a scalar α provides an equivalent description of the quantum state.
Moreover a superposition | ψ i = α| ψ1 i + β| ψ2 i of two state vectors
| ψ1 i and | ψ2 i describing two quantum states, also describes a new
possible quantum state. The new quantum state is also described by a
ray in Hilbert space and, therefore, the quantum state does not change
if the state vector is multiplied by an overall scalar. Therefore, we can
factorise α and write
β
| ψ i = α | ψ1 i + | ψ2 i , (2.14)
α
showing that the new quantum state is determined by the ratio β/α
rather than by an independent choice of α and β.
• Collapse of the state vector. The third principle states that as a result of
the measurement of the observable Ô, the state vector will change from
| ψ i to some eigenket | ωi i of Ô, with probability P (ωi ) = |h ωi | ψ i|2 ,
also corresponding to the probability of obtaining ωi as result of the
measurement (Born’s rule). This is equivalent to say that the process
of measurement has acted on the state describing the system as the
projection operator P̂ωi . Therefore, if we expand the state vector before
the measurement in the eigenbasis, writing
X
|ψi = | ωj ih ωj | ψ i , (2.15)
i
the measurement will act as a process inducing the collapse of the state
vector in a way that
| ψ i −→ | ψ 0 i = P̂ωi | ψ i = | ωi ih ωi | ψ i . (2.16)
2.3. HOW TO TEST QUANTUM MECHANICS 67
= h ψ |Ô| ψ i , (2.26)
where we have used, once more, the completeness relation. This expression is
very interesting since it shows that the expectation value can be calculated
without knowing the eigenvalues and eigenstates of Ô. Notice that if the
state of the system coincides with one eigenstate ωi , then simply hÔi = ωi .
Notice that we derived this result in the discrete case but the derivation can
be also easily extended to the continuous case.
We can also write an expression for the standard deviation, statistically
the average fluctuation around the mean value,
s
2 q
∆Ô = Ô − hÔi = hÔ2 i − hÔi2 , (2.27)
This implies that if one performs a third measurement, again of the observ-
able Ω̂, the Λ̂ eigenstate | λi i will now collapse in some eigenstate | ωi0 i that,
in general, does not coincide with | ωi i, explicitly:
A special situation occurs when the state state produced by the first measure-
ment, | ωi i, is unaffected by the second measurement: this can only happen
if | λi i is also an eigenstate of Ω̂, in a way to have a simultaneous eigenstate,
denoted by | ωi , λi i, of Ω̂ and Λ̂. If this happens, then one has at the same
time:
Ω̂| ωi , λi i = ωi | ωi , λi i , (2.33)
Λ̂| ωi , λi i = λi | ωi , λi i . (2.34)
2.5. COMPATIBLE AND INCOMPATIBLE OPERATORS 71
We can then operate again with Λ̂ in the first equation and with Ω̂ in the
second equation, obtaining the same result:
Λ̂ Ω̂| ωi , λi i = ωi λi | ωi , λi i , (2.35)
Ω̂ Λ̂| ωi , λi i = λi ωi | ωi , λi i , (2.36)
and subtracting the two equations side by side one obtains the condition:
[Ω̂, Λ̂] | ωi , λi i = 0 . (2.37)
This means that inverting the order of the two measurements the result will
be the same since in any case the measurement process on the state does not
alter the state itself.
We know already that two operators have common eigenbases when they
commute with each other. Therefore, for sure the condition [Ω̂, Λ̂] = 0
guarantees that they have simultaneous eigenkets and in the experiment we
discussed one would have simply P (ωi , λi , ωi0 ) = |h ωi | ψ i|2 if i0 = i and
P (ωi0 ) = 0 if i0 6= i. In this situation one says that the two operators are
compatible.
On the other hand if [Ω̂, Λ̂]| ψ i =
6 0 for any non trivial | ψ i, implying that
there is no common eigenket and any measurement of Λ̂ made on an eigenket
of Ω̂ will not give a well defined value (i.e., the uncertainty is alway non-
vanishing) and vice versa, then the two operators are said to be incompatible.
An important example is the case of position x̂ and momentum p̂ of an
elementary particle, since in this case, identifying p̂ = ~ k̂, where k̂ is the
conjugate operator of x̂, one has from Eq. (1.232)
[x̂, p̂] = i ~ . (2.38)
There is also a third case, when only some states are simultaneous eigen-
kets of both operators. In that case the condition (2.37) is satisfied only
for these states but not in general, for this reason it is not correct to write
[Ω̂, Λ̂] = 0, since the two operators do not commute with each other in gen-
eral. Notice that one cannot find a full basis of simultaneous eigenstates since
otherwise this would necessarily imply that the condition (2.37) is satisfied
for all states and this would fall within the case of compatible operators.
It is clear that if we invert the order of the two measurements we will obtain
exactly the same result, so that we can simply write:
The situation is quite simple in this non-degenerate case.4 The first mea-
surement induces the collapse of the quantum state, described by a ray in
Hilbert space, along one of the rays corresponding with one of the common
eigenstates of the two eigenstates with a probability given by the square of
the absolute value of the component of the initial state along that direction.
The collapse, as we have seen, is equivalent to project the initial state along
one of the eigenstates. Therefore, once the projection occurred, further mea-
surements of Ω̂ and/or Λ̂ will not change the quantum state, since as we
know P̂ωni = P̂ωi .
The degenerate case is just slightly more involved. A concrete exam-
ple helps understanding and it is quite easy to generalise. Consider the
three-dimensional real vector space (the Euclidean space). Suppose that the
operator Ω̂ has three different eigenvalues ω1 , ω2 , ω3 while a second opera-
tor Λ̂, commuting with Ω̂ and, therefore, compatible, has one eigenvalue λ3
with eigenket | ω3 , λ3 i also eigenket of Ω̂ with eigenvalue ω3 . This eigenket
identifies an axis in the Euclidean space (think of the z axis).
The other eigenvalue of Λ̂ is degenerate, so that λ = λ1 = λ2 but the
degeneracy is broken by Ω̂ and we can choose as basis of the eigenspace
4
If the operators are incompatible, then in general one has P (ωi , λi ) 6= P (λi , ωi ). Of
course in this case moreover the final quantum state does not have well-defined values
of ωi and λi , since after the second measurement one is in an eigenstate of the second
operator with the measured eigenvalue but this does not coincide with an eigenstate of
the first operator.
2.5. COMPATIBLE AND INCOMPATIBLE OPERATORS 73
| ψ i = α | ω3 , λ3 i + β | ω1 , λ i + γ | ω2 , λ i , (2.41)
If we now first measure Λ̂, the probability to obtain λ is given by the proba-
bility that initial state collapses on the λ subspace, and this is given by
P (λ) = β 2 + γ 2 . (2.43)
| P̂λ ψ i β| ω1 , λ i + γ | ω2 , λ i
| ψ0 i = = , (2.44)
h P̂λ ψ | P̂λ ψ i1/2 (β 2 + γ 2 )1/2
β2
P (λ, ω1 ) = (β 2 + γ 2 ) · = β2 , (2.45)
β2 + γ2
so that also in the degenerate case there is no difference in performing the
two measurements in different order. There is, however, a substantial differ-
ence: in the non-degenerate case the second measurement does not change
the quantum state obtained after the first measurement, so that the initial
74 CHAPTER 2. THE PRINCIPLES OF QUANTUM MECHANICS
quantum state directly collapses into the final quantum state, let us say
| ω1 , λ1 i. On the other hand, in the degenerate case, if one first measures
the degenerate eigenvalue the system goes through an intermediate quantum
state | ψ 0 i given by (2.44) that will then collapse into the final quantum state
after the second measurement. Of course for the special case γ = 0, this in-
termediate quantum state also coincides with the final one but, in general,
it does not.
Notice that in general, also Ω̂ might be degenerate and in this case one
could have a common eigenspace of quantum states corresponding to | ω, λ i
vectors. In this case one has to look for a third observable Γ̂ breaking the
degeneracy so that one can distinguish the states as | ω, λ, γi i, where the γi ’s
are the eigenvalue of Γ̂. Of course there might be necessity even of more
observable to fully break degeneracies and ultimately one obtains a set of
compatible operators providing a complete set of commuting observables.
time evolution
| ψ(t0 ) i −−−−−−−−−−−→ | ψ(t) i = Û (t, t0 ) | ψ(t0 ) i . (2.46)
As we have seen a generic state vector can be always expanded in the eigen-
basis {| ωi i} of some Hermitian operator Ω̂. Assuming a normalised vector,
the quantities |h ωi | ψ i|2 give the probabilities P (ωi ) to get the ωi ’s as result
of a measurement. The sum of all probabilities have to be equal to unity. At
the generic time t > t0 , this fundamental property has also to hold, so that
we can write X X
|h ωi | ψ(t0 i|2 = |h ωi | ψ(t i|2 = 1 . (2.47)
i i
2.6. THE SCHRÖDINGER EQUATION 75
It is easy then to see (PS n.7) that this straightforwardly translates into the
result that the time-evolution operator has to be unitary:
Notice that this also implies that the state vector remains normalised at all
later times. Notice that if the state first evolve from t0 to some time t1 and
then from t1 to some time t2 , i.e.,
If we now consider the case when the state vector evolves from t0 to a time
t0 + dt, where dt is an infinitesimal interval of time, having assumed that
time is a continuous variable, the unitarity operator has also to satisfy the
property
lim Û (t0 + dt, t0 ) = Iˆ . (2.51)
dt→0
These three properties, unitarity, composition and continuity, imply that the
time-evolution operator for such infinitesimal time displacement has to be
necessarily of the form (see PS n.7)
Ĥ
Û (t0 + dt, t0 ) = Iˆ − i dt , (2.52)
~
Ĥ
Û (t + dt, t0 ) − Û (t, t0 ) = −i dt Û (t, t0 ) , (2.54)
~
that, in differential form, gives the Schrödinger equation for the time-evolution
operator
∂
i~ Û (t, t0 ) = Ĥ Û (t, t0 ) . (2.55)
∂t
When this is applied to the state vector | ψ(t0 ) i, one obtains the Schrödinger
equation for the time evolution of the state vector:
∂
i~ | ψ(t) i = Ĥ| ψ(t) i . (2.56)
∂t
The Schrödinger equation is a first order differential equation in time and,
therefore, it is sufficient to know the initial state vector to determine its time
evolution.
Since the | Ei i’s are linearly independent, then necessarily each component
has to satisfy
dai (t) i Ei
=− ai (t) , (2.61)
dt ~
6
For definiteness we write we are referring to the discrete case but all expressions can
be extended to the continuous case as usual.
78 CHAPTER 2. THE PRINCIPLES OF QUANTUM MECHANICS
with solution i Ei
ai (t) = ai (0) e− ~
t
, (2.62)
where we have taken t0 = 0. The solution of the Schrödinger equation is
then given by X i Ei
| ψ(t) i = ai (0) e− ~ t | Ei i . (2.63)
i
It is remarkable that if the initial state coincides with one of the energy
eigenstates, let us say | ψ(0) i = | Ej i, then
i Ej
| ψ(t) i = e− ~
t
| Ej i , (2.64)
meaning that the system will remain in the same initial quantum state. It is
also easy to see that in this case the probability distribution P (ωi ) for any
generic observable Ω̂ is time-independent:
= |h ωi | Ej i|2
= P (ωi , 0) .
This implies that in this case the expectation values and the standard devi-
ations of any generic observable Ω̂ is independent of time. For these reasons
the energy eigenstates are also called stationary states.
with time. We can now indeed verify, as an exercise, that the Schrödinger
equation preserves the norm of the state. This can be proved quite easily
directly calculating the time derivative of a state obeying the Schrödinger
equation. Let us then start writing:
d ∂ψ(t) ∂ψ(t)
h ψ(t) | ψ(t) i = hψ(t) | i+h | ψ(t) i , (2.66)
dt ∂t ∂t
where | ∂ψ(t)/∂t i ≡ ∂| ψ(t) i/∂t and h ∂ψ(t)/∂t | is its adjoint.7 The Schrödinger
equation can be recast as
∂ψ(t) i
= − Ĥ | ψ(t) i . (2.67)
∂t ~
When these equations are inserted into Eq. (2.66), one immediately finds, as
expected, that the right-hand side vanishes, so that
d
h ψ(t) | ψ(t) i = 0 . (2.69)
dt
In particular, if the state is initially normalised to unity, it will remain nor-
malised to unity. As we said this result was expected, since it trivially follows
from the unitarity of the time-evolution operator. However, it is a useful pre-
liminary exercise for the proof of a more general result.
and using again the Eqs.(2.67) and (2.68), one immediately finds Ehrenfest’s
theorem:
dhQ̂i i Dh iE
=− Q̂, Ĥ . (2.71)
dt ~
This equation is interesting, since it says that if the operator commutes
with the energy, then its expectation value is conserved. It is clear that if
Q̂ = Ĥ itself, then one finds immediately dhĤi/dt = 0. This shows that for
a time-independent Hamiltonian its expectation value, that can be identified
with the energy of the system, is a constant of motion: this result is clearly
the classical analogue of energy conservation for an isolated system. More
generally, it is interesting that from Ehrenfest’s theorem one finds that the
laws of classical mechanics are recovered for the expectation values of the
QM operators. Finally, notice that the conservation of the norm of the state
vector can be regarded as a special case of Ehrenfest’s theorem for the simple
ˆ
choice Q̂ = I.
and we know that at the time t the state vector evolves into ψ(t) given by
Eq. (2.63). Given an observable Q̂, we can then calculate the time evolution
of its expectation value,
In order to highlight the main feature of the time-evolution of hQ̂i, let us con-
sider the simple case when n = 2, this can be easily generalised to arbitrary
n. We have then
E1 E2
| ψ(t) i = e−i ~
t
a1 (0) | E1 i + e−i ~
t
a2 (0) | E2 i , (2.73)
2.6. THE SCHRÖDINGER EQUATION 81
where Qij ≡ h Ei |Q̂| Ej i are of course the matrix elements of Q̂ in the energy
eigenbasis and we used Q21 = Q?12 . The last two terms in the right-hand side
are complex conjugate of each other so that we can also write
h (E2 −E1 )
i
hQ̂i(t) = |a1 (0)|2 Q11 + |a2 (0)|2 Q22 + 2Re a?1 (0) a2 (0) Q12 e−i ~ t .
(2.75)
Considering that given two complex numbers z1 and z2 one has Re(z1 z2 ) =
Re(z1 ) Re(z2 )−Im(z1 ) Im(z2 ), we can rewrite the last term in a way to obtain
hQ̂i(t) = |a1 (0)|2 Q11 + |a2 (0)|2 Q22 + 2Re [a?1 (0) a2 (0) Q12 ] cos α(t)
−2Im [a?1 (0) a2 (0) Q12 ] sin α(t) , (2.76)
where we defined
E2 − E1
α(t) ≡ t. (2.77)
~
Notice that at the time t = 0 one has
hQ̂i(0) = |a1 (0)|2 Q11 + |a2 (0)|2 Q22 + 2Re [a?1 (0) a2 (0) Q12 ] . (2.78)
With some simple algebraic steps, one can then re-express hQ̂i(t) as
α(t)
hQ̂i(t) = hQ̂i(0)−4Re[a?1 (0) a2 (0) Q12 ] sin2 −2Im[a?1 (0) a2 (0) Q12 ] sin α(t) ,
2
(2.79)
2
where we used the simple trigonometric formula cos α(t)−1 = −2 sin [α(t)/2].
As one can see the expression exhibits an oscillatory behaviour around the
initial value. It simplifies even further if a1 (0), a2 (0) and Q12 are real, since
in this case the last term on the right-hand side containing the imaginary
part vanishes and simply
α(t)
hQ̂i(t) = hQ̂i(0) − 4 a1 (0)a2 (0)Q12 sin2 . (2.80)
2
82 CHAPTER 2. THE PRINCIPLES OF QUANTUM MECHANICS
Using the expression (2.80) and making the proper identifications, one finds
(see PS. 8) that the probability of an electron neutrino to be produced in the
centre of the Sun and be detected as either a muon or tauon neutrino on the
Earth is given by9
c4 ∆m212
2 2
Pνe →νµ+τ (t) = sin (2θ12 ) sin t , (2.86)
4cp~
where we defined ∆m212 ≡ m22 − m12 . The time t is the time that takes
to neutrinos to travel to us from the Sun and it is approximately given by
t ' `/c, where ` ' 1 A.U. is the baseline length and is of course given,
in the case of solar neutrinos, by the Sun-Earth distance. This formula,
with a proper replacement of the mixing angle, mass squared difference and
baseline length, applies also to a few other neutrino oscillation experimental
setups, in particular to atmospheric neutrinos. In this case muon neutrinos
are produced in the outer layer of the atmosphere and can be detected as
tauon neutrinos on the Earth surface.
This expression shows that the time-evolution operator, in the case of time-
independent Hamiltonian operator, is given by
X Ei
Û (t) = e−i ~
t
| Ei i h Ei | . (2.88)
i
9
Notice that the probability that the neutrino state is detected as an electron neutrino
state, like at the production, is of course simply given by
and is the so-called the (electron neutrino) survival probability, while Pνe →νµ+τ (t) is the
so-called (muon-tauon neutrino) appearance probability.
84 CHAPTER 2. THE PRINCIPLES OF QUANTUM MECHANICS
As an exercise, one can verify that this expression for Û (t) respects unitarity:
X (Ej −Ei )
Û † (t) Û (t) = | Ei i h Ei | Ej i h Ej | e−i ~ t (2.89)
i,j
X
= | Ei i h Ei |
i
= Iˆ .
We have so far assumed that the state vector describing the system also
describes its time evolution. However, there is an alternative description of
time evolution in QM considering that all physical observables are encoded
in matrix elements of operators of the kind h ψi (t) |Ô| ψj (t) i. If we redefine
state vectors and operators according to the following transformations
| ψ(t) iH = Û † | ψ(t) iS = | ψ(0) iS , (2.90)
ÔH (t) = Û † (t) ÔS Û (t) , (2.91)
where we denoted by the subscript ‘S’ quantities in the Schrödinger pic-
ture that we discussed so far, and with the subscript ‘H’ quantities in the
Heisenberg picture, it is straightforward to check that matrix elements are
invariant:
H h ψi |ÔH (t)| ψj iH = S h ψi (t) |ÔS (t)| ψj (t) iS . (2.92)
In this way, while in the Schrödinger picture state vectors evolve with time
and operators are usually time-independent, in the Heisenberg picture state
vectors are time independent and all operators are necessarily time depen-
dent. Notice that these two simultaneous unitary transformations of states
and operators, bringing from the Schrödinger to the Heisenberg picture,
are a particular example of the general double transformation performed in
Eq (1.139) combining an active (transforming states) and a passive (trans-
forming operators) transformation that leave invariant matrix elements. In
the following we will continue to work in the Schrödinger picture dropping
the subscript ‘S’.
We can first of all take the inner products of both sides of (2.56) with an
x̂ basis eigenbra h x |, obtaining
d
i~ h x | ψ(t) i = h x |Ĥ| ψ(t) i . (2.93)
dt
Similarly to the results obtained for the matrix elements of x̂ and k̂, respec-
tively Eq. (1.193) and Eq. (1.214), the matrix elements of Ĥ in the x̂-basis
can be written as
so that one immediately obtains the Schrödinger equation for the wave func-
tion in position basis:
∂ψ(x, t)
i~ = H(x, p(x))ψ(x, t) . (2.96)
∂t
86 CHAPTER 2. THE PRINCIPLES OF QUANTUM MECHANICS
Bibliography
87
88 BIBLIOGRAPHY
Chapter 3
Harmonic Oscillator in
Quantum Mechanics
p2 1
H(x, p) = + m ω 2 x2 , (3.1)
2m 2
89
90CHAPTER 3. HARMONIC OSCILLATOR IN QUANTUM MECHANICS
p̂2 1
Ĥ(x̂, p̂) = + m ω x̂2 . (3.2)
2m 2
It is convenient to introduce the dimensionless operators1
r r
mω 1
X̂ = x̂ and P̂ = p̂ . (3.3)
~ m~ω
We have seen that this commutation relation for position and momentum
operators can be derived by identifying p̂ = ~k̂ = −i ~D̂, where k̂ is the
conjugate of x̂. Is this identification to be meant as an additional postulate
to the four we enunciated? This is one possibility, the other possibility is
to assume the commutation relation (3.6) as a postulate and in this case
one can show that from this one can derive p̂ = ~ k̂. However, even more
fundamentally, one can show that the identification p̂ = ~k̂ is a consequence
of identifying momentum, as in classical mechanics, with the generator of
translations and, if space is homogeneous its averaged value is a conserved
quantity of motion, analogously to what happens in classical mechanics where
from Nöther theorem one derives that momentum is a conserved quantity.
From this identification of momentum with the generator of translation it
follows, in our one-dimensional case, then necessarily p̂ = C D̂. The constant
of proportionality C needs to be purely imaginary in a way for the momentum
to be Hermitian, so that we can write p̂ = −i ~ D̂ where ~ is the Planck
constant and has the dimensionality of angular momentum. The Planck
1
Also called scaled position operator and scaled momentum operator respectively.
3.2. LADDER OPERATORS 91
1
â = √ X̂ + i P̂ (3.7)
2
Therefore, one has ↠â 6= â ↠and explicitly one can write
† 1
â â = X̂ + i P̂ (X̂ − i P̂ ) (3.10)
2
1
2 2
1
= X̂ + P̂ − i [X̂, P̂ ] (3.11)
2 2
Ĥ 1
= + (3.12)
~ω 2
2
The hat on these operators is often omitted, especially in historical papers and text-
books. We are here keeping it, following more recent literature, to be utterly unambiguous.
92CHAPTER 3. HARMONIC OSCILLATOR IN QUANTUM MECHANICS
and
1
↠â =
X̂ − i P̂ (X̂ + i P̂ ) (3.13)
2
1 2 1
= X̂ + P̂ 2 + i [X̂, P̂ ] (3.14)
2 2
Ĥ 1
= − . (3.15)
~ω 2
Subtracting the two, one can verify again that â, ↠= 1. In terms of
the dimensionless number operator N̂ ≡ ↠â, the Hamiltonian can then be
expressed as3
1
Ĥ = ~ ω N̂ + . (3.16)
2
The Hamiltonian eigenstates now can be labelled in terms of eigenvalues of
N̂ denoted by n, so that the eigenvalue equation for N̂ can be written as
N̂ | n i = n | n i , (3.17)
where the eigenstates | n i are normalised to unity, i.e., h n | n i = 1. The
eigenstate | n i is clearly also an eigenstate of Ĥ with eigenvalue En = (n +
1/2) ~ω. The eigenvalues n of N̂ have to be necessarily not only real, since
N̂ is Hermitian, but also nonnegative since
n = h n |N̂ | n i = h n |↠â| n i = h â n | â n i = ||â n||2 ≥ 0 . (3.18)
We want now to understand the action of the ladder operators on the eigen-
states | n i and the result will justify their names. In this respect, it is useful
to notice that
[N̂ , ↠] = [↠â, ↠] (3.19)
= ↠â ↠− ↠↠â
= ↠[â, ↠]
= â†
and
[N̂ , â] = [↠â, â] (3.20)
= ↠â â − â ↠â
= [↠, â] â
= − â .
3
Notice that N̂ is also often found to be written without the hat symbol in the literature.
3.2. LADDER OPERATORS 93
With these results we can now easily solve the eigenvalue problem for N̂
understanding the action of â and ↠. If we first act with the number operator
on ↠| n i, we find
† † †
N̂ â | n i = â N̂ + [N̂ , â ] | n i (3.21)
= ↠(N̂ + 1) | n i (3.22)
= (n + 1) ↠| n i , (3.23)
Notice that this result now justifies why ↠is referred to as the raising oper-
ator.4
We can repeat the same procedure applying now the number operator to
the kets â | n i, finding:
N̂ â | n i = â N̂ + [N̂ , â] | n i (3.26)
= â (N̂ − 1) | n i (3.27)
= (n − 1) â | n i . (3.28)
â | n i = Cn− | n − 1 i , (3.29)
1
Vn ≡ h n |V̂ | n i = ~ω h n |X̂ 2 | n i . (3.35)
2
5
It is also called the destruction operator, since in quantum field theory an elementary
particle of a given type is destroyed by applying the associated destruction operator to an
initial quantum state.
6
Mnemonical note: the larger quantum number labelling the kets appears in the square
root.
3.3. MATRIX REPRESENTATIONS 95
From the definition of lowering and raising operators Eqs. (3.7) and (3.8),
one can simply write
â + â†
X̂ = √ (3.36)
2
so that
2 â â + â ↠+ ↠â + ↠â†
X̂ = . (3.37)
2
Using the orthonormality of the eigenstates | n i, one has simply h n |↠↠| n i =
h n |â â| n i = 0. Using then â ↠= [â, ↠] + ↠â = 1 + N̂ , one easily obtains
1 1 1
Vn = ~ω n + = En . (3.38)
2 2 2
Notice the difference between the null vector 0 and the ground state | 0 i.
The easiest matrix representation is clearly the one of the number opera-
tor N̂ since its matrix elements Nn0 n = h n0 |N̂ | n i = n δn0 n so that one simply
has:
0 0 0 0 ...
0 1 0 0
0 0 2 0
N̂ → . (3.40)
0 0 0 3
.. ..
. .
Clearly the matrix representing Ĥ is also diagonal, with energy levels En =
~ ω (n + 1/2) on the diagonal.
The matrix elements of the raising operator are given by (n0 , n = 0, 1, 2, . . . )
√ √
a†n0 n = h n0 |↠| n i = n + 1h n0 | n + 1 i = n + 1 δn0 ,n+1 , (3.41)
96CHAPTER 3. HARMONIC OSCILLATOR IN QUANTUM MECHANICS
in a way that one has (notice that the first row and the first column are the
zero-th row and the zero-th column)
0 0 0 0 ...
11/2 0 0 0 ...
† 0 21/2 0 0 ...
â → . (3.42)
0
0 31/2 0 ...
.. ...
.
Analogously the matrix elements of the lowering operator are given by (n0 , n =
0, 1, 2, . . . )
√ √
an0 n = h n0 |â| n i = nh n0 | n − 1 i = n δn0 ,n−1 . (3.43)
ψn (x) ≡ h x | n i . (3.45)
We can now calculate the ground state wave function ψ0 (X) considering
that, since â| 0 i = 0, then necessarily h X |â| 0 i = 0. This means that the
wave function of the null vector vanishes for all values of X. From the
definition of â (see Eq. (3.7)), we can then recast this simple result as
From the representations of X̂ and P̂ in the X̂ basis, we than find the simple
differential equation for the ground state wave function
∂
X ψ0 (X) + ψ0 (X) = 0 , (3.48)
∂X
with solution
1 2
ψ0 (X) = C e− 2 X , (3.49)
where C is an arbitrary constant. In the x̂-basis, this simply translates into
1 m ω x2
ψ0 (x) = c e− 2 ~ω . (3.50)
The constants C and c, as usual, can be conveniently chosen real and fixed
by normalising ψ0 (x) to unity, i.e., imposing
Z +∞
dX ψ0? (X) ψ0 (X) = 1 , (3.51)
−∞
finding
14
mω 2
− 41
C=π and c = . (3.52)
π~ω
We can now calculate the normalised wave function of the first excited state
| 1 i = ↠| 0 i in the X̂-basis:
99
100 BIBLIOGRAPHY
Chapter 4
Angular momentum
4.1 Definition
Angular momentum can be defined in classical mechanics as the generator of
rotations and, for an isolated system, it is a conserved quantity of motion if
one assumes isotropy of space.1 Analogously to what we have seen in the case
of the momentum, one can then generalise the same definition to QM saying
that the angular momentum vector operator is the generator of rotations
(times the Planck constant ~). This definition is analogous to the definition
of Hamiltonian operator as generator of time-evolution. Let us consider a
rotation of the physical system by an angle θ around an axis along the unit
vector n. This will induce a transformation of the state vector | ψ i described
by an operator of rotation D̂(n, θ) 2 into a rotated state vector | ψ 0 (n, θ) i,
explicitly:
D̂(n,θ)
| ψ i −−−−→ | ψ 0 (n, θ) i = D̂(n, θ) | ψ i . (4.1)
1
More precisely, the angular momentum L · θ is the generator of a rotation of an angle
θ about an axis parallel to the unit vector n.
2
This is a generalization of the rotation operator R̂n (θ) acting on vectors in the Eu-
clidean space seen in problem sheets.
101
102 CHAPTER 4. ANGULAR MOMENTUM
Ĵ · n
D̂(n, dθ) = Iˆ − i dθ . (4.2)
~
The vector operator
Ĵ = Jˆx i + Jˆy j + Jˆz k (4.3)
is the generator of rotations and is called angular momentum (vector) op-
erator. Notice that, in the same way as the unitarity of the time-evolution
operator implies that the Hamiltonian operator has to be Hermitian, now
the unitarity of the rotation operator implies that Ĵ has to be Hermitian.
The crucial difference with classical mechanics is that in the case of QM
the angular momentum is the sum of two components, namely
Ĵ = L̂ + Ŝ . (4.4)
• The spin angular momentum operator Ŝ does not have a classical cor-
respondence.
Defining Jˆ1 = Jˆx , Jˆ2 = Jˆy , Jˆ3 = Jˆz , from the general definition of angular mo-
mentum as generator of rotations, one can derive the following fundamental
commutation relations of angular momentum (i, j, k = 1, 2, 3):
h i
Jˆi , Jˆj = i ~ εijk Jˆk , (4.5)
(ii) It is fully antisymmetric under the exchange of any two indexes (for
example: εjik = −εijk ).
This implies that all 21 entries where at least two indexes are equal vanish,
while ε123 = ε312 = ε231 = 1 and ε213 = ε321 = ε132 = −1.
For example, one has [Jˆx , Jˆy ] = i ~ Jˆz . We will not prove the relations
Eqs. (4.5) in full generality, but we will see that they hold in the specific case
of the orbital angular momentum of a particle.
4.2. LADDER OPERATORS FOR ANGULAR MOMENTUM 103
This implies that one can always find a common eigenbasis for Jˆ2 and one of
Jˆk , conventionally Jˆz . If we indicate with λ ~2 the eigenvalues of Jˆ2 and with
m ~ the eigenvalues of Jˆz , then their common normalised eigenstates can be
denoted by | λ, m i and, therefore, we can write:
At this stage λ and m can be any real numbers but we now show that the
fundamental commutation relations enforce stringent limitations. First of all
we can introduce the ladder operators (for angular momentum) defined as:
where notice that Jˆ±† = Jˆ∓ . They satisfy the commutation relations
h i
Jˆz , Jˆ± = ±~ Jˆ± , [Jˆ+ , Jˆ− ] = 2~ Jˆz and [Jˆ2 , Jˆ± ] = 0 , (4.12)
104 CHAPTER 4. ANGULAR MOMENTUM
±
Choosing conventionally Cλm real and positive, i.e., setting a phase factor
±
p
equal to unity, we find Cλm = λ − m(m ± 1) so that we can finally write
Jˆ− Jˆ+ = Jˆ2 − Jˆz2 − ~ Jˆz and Jˆ+ Jˆ− = Jˆ2 − Jˆz2 + ~ Jˆz , (4.20)
showing that
1ˆ ˆ
Jˆ2 − Jˆz2 = J− J+ + Jˆ+ Jˆ− . (4.21)
2
Using that Jˆ− and Jˆ+ are the adjoint of each other, one has
1
h λ, m |Jˆ2 − Jˆz2 | λ, m i = h λ, m | Jˆ− Jˆ+ + Jˆ+ Jˆ− | λ, m i (4.22)
2
1 ˆ
= h J+ λm | Jˆ+ λm i + h Jˆ− λm | Jˆ− λm i ≥ 0 ,
2
where, as usual, we defined | Jˆ± λm i ≡ Jˆ± | λ, m i. At the same time, from
the eigenvalue equations (4.10), one obtains
0 ≤ m2 ≤ λ , (4.24)
showing that λ ≥ 0 and that m2 has an upper bound, implying that there
exists both a maximum value mmax , such that m ≤ mmax , and a minimum
value mmin , such that m ≥ mmin . This implies both
and
Jˆ− | λ, mmin i = 0 . (4.26)
Let us start to see the implications of the first of these two conditions,
Eq. (4.25). It also implies
Similarly, we can proceed from the condition in Eq. (4.114) and write
A comparison of Eq. (4.29) with Eq. (4.32), immediately shows that mmin =
−mmax . The value mmax is usually simply denoted by j, in a way that
Eq. (4.29) can also be written as λ = j (j + 1) and one has
−j ≤ m ≤ j . (4.33)
Notice that j 2 < λ = j(j + 1). This means that the absolute value of the
angular momentum along the z direction can never be equal to the total
angular momentum, as in classical mechanics, but is always lower. This is
because if j 2 = λ, then the x and y components would vanish but in such a
situation all components of the angular momentum would be simultaneously
determined and this would contradict the commutation relations.
Clearly, applying successively Jˆ+ to | λ, −j i, one obtains | λ, j i in n = 2 j
steps, where n is some integer. Therefore, one obtains j = n/2, meaning that
j can be either an integer number, if n is even, or a half-integer number, if n
is odd. This implies that if j is an integer, then all values of m are integers
and if j is half-integer, then all values of m are half-integers. For a given j,
one has then 2j + 1 allowed values of m given by:
m = −j, −j + 1, . . . , j − 1, j . (4.34)
• j = 0 ⇒ m = 0;
• j = 1 ⇒ m = −1, 0, +1;
Jˆ2 | j, m i = j (j + 1) ~2 | j, m i , (4.35)
Jˆz | j, m i = m~ | j, m i . (4.36)
It should be noticed how all results have been obtained in great generality just
from the fundamental commutation relations (4.5) and these just follow from
the general definition of angular momentum vector operator as the generator
of rotations.
We can finally draw our attention to the ladder operators. First of all let us
rewrite Eq. (4.19) replacing λ → j(j + 1):
Let us now consider the operators Jˆx and Jˆy . These are straightforwardly
expressed in terms of the ladder operators as:
We can now make a useful exercise solving the eigenvalue problem for Jˆx
(analogous results can be easily obtained for Jˆy ). First of all one can easily
check that Jˆx (and likewise Jˆy ) has the same eigenvalues as Jˆz , i.e., ~, 0 and
−~. This is expected from space isotropy, since all directions are equivalent.
4.5. ORBITAL ANGULAR MOMENTUM 109
Analogously, one finds P (mx = 0) = 1/2 and P (mx = −1) = 1/4. After the
measurement of the x component one could again measure the z component
of the angular momentum. This time the initial state is in one of the L̂x eigen-
states. One can now again calculate the probability to obtain as outcome one
of the three eigenvalues of L̂z . Suppose that the outcome of the x component
measurement was mx = 1, this time one has P (m) = |h 1, m | 1, 1x i|2 . For
example the probability to obtain m = 1 would be P (1) = 1/4 and not 1 as
one could naively expect.
L̂ = r̂ × p̂ (4.48)
3
Notice that using the Levi-Civita Ptensor one can also write the components of a generic
vectorial product ~c = ~a × ~b as ci = j,k εijk aj bk .
110 CHAPTER 4. ANGULAR MOMENTUM
(where x̂1 = x̂, x̂2 = ŷ, x̂3 = ẑ and correspondingly for the three momenta).
This is a compact expression implying that the commutators for the same
components are given by
[x̂i , p̂i ] = i ~ , (4.53)
while the commutators [x̂i , p̂j ] = 0 for i 6= j. These commutation rela-
tions can be easily derived similarly to the one-dimensional case. From these
commutation relations for positions and momenta, one can then derive the
commutation relations for the operators of the orbital angular momentum
components: h i
L̂i , L̂j = i ~ εijk L̂k , (4.54)
that, when spin is also included, are generalised by the Eqs. (4.5), as we
pointed out.
Proof. Let us prove the commutation relations (4.54) in the case i = 1
and j = 2. Using the commutation relations (4.52), one has:
Similar derivations can be repeated for the other two non-vanishing commu-
tators, confirming Eq. (4.54).
4.5. ORBITAL ANGULAR MOMENTUM 111
We are now interested to see how the orbital angular momentum oper-
ator acts on on the wave function in the position basis, this time in three
dimensions, that of course will be given by
ψ(~r) ≡ h ~r | ψ i . (4.56)
to fully determine the wave function: they are not a complete set of opera-
tors. This is because one needs another operator, usually the Hamiltonian,
to determine also the radial part. We can denote generically by n the addi-
tional quantum number that is needed to characterise the radial part so that
the full quantum states can be denoted by kets | n, l, m i and the associated
wave functions will be given by ψnlm (~r) = h ~r | n, l, m i. If the system has a
spherical symmetry. This means that the particle moves in a potential with
spherical symmetry that depends only on r but not on θ and φ, then the
solutions will factorise in a radial part Rnl (r) and in an angular part given
by the spherical harmonics Ylm (θ, φ), explicitly:
∂
h r, θ, φ |L̂z | n, l, m i = −i ~ Rnl (r) Ylm (θ, φ) . (4.67)
∂φ
One then obtains a differential equation from which the radial part drops
out:
∂
Ylm (θ, φ) = i m Ylm (θ, φ) . (4.69)
∂φ
This has as a solution:
4.6 Spin
In this section we now discuss the spin angular momentum, first its general
properties and then some specific applications.
of the total angular momentum vector operator Ĵ (see Eqs. (4.5)), then nec-
essarily the same commutation relations have to be also satisfied separately
by the components of the orbital angular momentum vector operator L̂ and
by the components of the spin Ŝ that is also a vector operator and we can,
therefore, write: h i
Ŝi , Ŝj = i ~ εijk Ŝk . (4.71)
In the case of the orbital angular momentum, we saw that these are consistent
with its definition coming from the classical correspondent quantity and from
the canonical commutation relations (4.52). In the case of spin there is no
such classical analogy but the same commutation relations need to hold since
it is a component of the total angular momentum and, in the case that the
orbital angular momentum vanishes, like in the Stern-Gerlach experiment, it
coincides with it.
The quantum numbers of spin are denoted ms , for the z component Ŝz ,
and its maximum value by s, while the corresponding eigenvalue of the square
of the spin operator, Ŝ 2 , is given by ~2 s (s + 1). Of course all results we ob-
tained for the angular momentum Ĵ also apply to Ŝ, that is a specific case
when orbital angular momentum can be neglected. This is because spin obeys
the same commutation relations and, as we have seen, all results we obtained
for angular momentum were derived just from the fundamental commutation
relations (4.5). In particular, for each value of s, there are 2s + 1 possible
values of ms . The clear difference between orbital angular momentum and
spin is that while in the case of spin both half-integer and integer values are
possible, in the case of orbital angular momentum only integer values are
possible, as we discussed. Moreover, spin does not have a classical counter-
part and its quantum numbers are not related to any kinematic or dynamical
variable describing the motion of the system, they are indeed an example of
internal quantum numbers).
Spin operators act on spin state vectors. They reside in a vector space
Vs that is distinct from the Hilbert space Vo describing the orbital degrees
of freedom of the system, related to its motion and position in space. The
Hilbert space Vo is infinite-dimensional and we can denote the generic state
vector by ψo . The eigenstates can be denoted by | n, l, m i, where l and m are
the quantum numbers of angular momentum and n are some quantum num-
bers describing other dynamical quantities of the system such as energy. The
vector space Vs has dimension 2s + 1 and it is, therefore, finite-dimensional
and we can generically denote a state vector in Vs by | ψs i. The full Hilbert
space of the system will be then given by the direct product (also called tensor
4.6. SPIN 115
product or Kronecker product) of spin and orbital vector spaces and denoted
by
V = Vo ⊗ Vs . (4.72)
If the orbital and spin degrees of freedom evolve independently of each other,
and this happens if the Hamiltonian operator is separable and can be written
as the sum of an orbital contribution Ĥo and a spin contribution Ĥs , explicitly
Ĥ = Ĥs + Ĥo , (4.73)
then consequently a generic state vector | ψ i factorises into
| ψ i = | ψo (t) i ⊗ | ψs (t) i , (4.74)
where the evolution of | ψo i is described by Ĥo and the evolution of | ψs i is
described by Ĥs .
confirming, also in this special case, the general result that Jˆ2 is propor-
tional to the identity operator (see Eq. (4.37)). Therefore, its action on a
generic state vector | ψ i is simply given by Ŝ 2 | ψ i = (3/4)~2 | ψ i and the
wave function transforms accordingly:
+ +
2 3 2 1 0 ψs 3 2 ψs 3
h ~r |Ŝ | ψ i = ~ ψo (~r) − = ~ ψo (~r) − = ~2 ψ(~r) .
4 0 1 ψs 4 ψs 4
(4.79)
One can also easily find the matrix representations for the spin com-
ponents that can be conveniently expressed in terms of the Pauli matrices
~σ = (σx , σy , σz ) as
~ = ~ ~σ ,
S (4.80)
2
4.6. SPIN 117
where:
0 1 0 −i 1 0
σx = , σy = , σz = . (4.81)
1 0 i 0 0 −1
However, it is easy to check that the Pauli matrices also satisfy the anti-
commutation relations
{σi , σj } ≡ σi σj + σj σi = 0 , (4.83)
σi σj = i εijk σk . (4.84)
We can now solve the eigenvalue problems for the spin components. For all
three spin components the eigenvalues, as expected, are the same: +~/2 and
−~/2. One then finds the following eigenvectors:
• For Ŝx :
1 1 1 1 1 1 1 1
| , i ≡ | +x i → √ and | , − ix ≡ | −x i → √ ;
2 2x 2 1 2 2 2 −1
(4.86)
These results can be also generalised to the case of the spin component Ŝn̂
along an arbitrary direction n̂ = (θ, φ). This can be expressed as:
where in Cartesian coordinates n̂ = (sin θ cos φ, sin θ sin φ, cos θ). Also in this
case the eigenvalues are ±~/2 (they are invariant under a rotation!). Using
the useful short notation already adopted for the Cartesian components,
1 1
| ±n̂ i ≡ | , ± i , (4.89)
2 2 n̂
we can then write the eigenvalue equation as
~
Ŝn̂ | ±n̂ i = ± | ±n̂ i , (4.90)
2
finding for the eigenvectors the following representations in the Ŝz -basis (see
problemsheets):
cos(θ/2)
| +n̂ i → , (4.91)
sin(θ/2) e+iφ
sin(θ/2)
| −n̂ i → . (4.92)
− cos(θ/2) e+iφ
These expressions show clearly that a spinor is not single-valued, since per-
forming a 2π rotation around the axis n̂, the spinor changes sign.
Finally, from the general results found for the generic angular momentum
operators and in particular for the matrix elements of the ladder operators of
angular momentum in Eqs. (4.19), one can easily find for the representations
of the spin ladder operators (j = s = 1/2):
0 1 0 0
S+ = ~ and S− = ~ . (4.93)
0 0 1 0
(1) (2)
[Jˆi , Jˆj ] = 0 ∀ i, j = 1, 2, 3 . (4.94)
On the other hand, if one takes two components of the same angular momen-
tum, then they satisfy the fundamental commutation relations (4.5) that get
specialised into
(1) (1) (1) (2) (2) (2)
[Jˆi , Jˆj ] = i~ εijk Jˆk , and [Jˆi , Jˆj ] = i~ εijk Jˆk . (4.95)
Its components also satisfy the fundamental commutation relations (4.5), i.e.,
(1) (2) (1) (2)
X (1) (2)
[Jˆi + Jˆi , Jˆj + Jˆj ] = i ~ εijk (Jˆk + Jˆk ) , (4.98)
k
called the uncoupled basis) and to the second as the total-j basis (sometimes
also called the coupled basis).
The total-j basis is of course a simultaneous eigenbasis of the four op-
(1)
erators Jˆ2 , Jˆz , (Jˆ(1) )2 , (Jˆ(2) )2 , where essentially Jˆ2 and Jˆz replace Jz and
(2)
Jz . This is a very useful change of basis in order to solve problems where
the Hamiltonian operator depends on terms where the two angular momenta
couple together, such as terms ∝ Ĵ(1) · Ĵ(2) .
Notice that all four operators Jˆ2 , Jˆz , (Jˆ(1) )2 , (Jˆ(2) )2 do commute with each
other and, therefore, they admit a common eigenbasis. We already know
that [Jˆ2 , Jˆz ] = [J (1) )2 , (Jˆ(2) )2 ] = 0. We are then left to show that also
[Jˆ2 , (Jˆ(1) )2 ] = [Jˆ2 , (Jˆ(2) )2 ] = 0. First of all notice that we can write:
Jˆ2 = Ĵ· Ĵ = (Ĵ(1) + Ĵ(2) )·(Ĵ(1) + Ĵ(2) ) = (Jˆ(1) )2 +(Jˆ(2) )2 +2 Ĵ(1) · Ĵ(2) , (4.99)
where in the last terms we used the fact that Ĵ(1) · Ĵ(2) = Ĵ(2) · Ĵ(1) . Notice
now that Jˆ2 does indeed commute with each of the three terms on the right-
hand side and notice also that this expression clearly shows the convenience
of using the total-j basis instead of the product basis when the Hamiltonian
contains terms proportional to Ĵ(1) · Ĵ(2) .
Therefore, the common eigenvectors | j m, j1 j2 i also form a basis for the
states of angular momentum of the combined system. Let us explicitly write
the eigenvalue equation for each operator:
If one fixes the values of j1 and j2 , in common with the two bases, this
defines a subspace of V1⊗2 . Since j1 and j2 are fixed, one usually simplifies
the notation writing:
| j, m i ≡ | jm, j1 j2 i , (4.101)
| m1 , m2 i ≡ | j1 m1 , j2 m2 i . (4.102)
It is easy to calculate the dimension of the subspace V1⊗2 counting the number
of distinct eigenvectors | m1 , m2 i. Since m1 and m2 can take (2j1 + 1) and
(2j2 + 1) values respectively, this is clearly given by (2j1 + 1) (2j2 + 1). This
has necessarily to coincide with the number of distinct eigenvectors | j, m i.
4.8. CLEBSCH-GORDAN COEFFICIENTS 121
(1) (2)
First of all since Jˆz = Jˆz + Jˆz , then each value of m has to be simply
related to m1 and m2 by m = m1 +m2 .6 The maximum value of j, denoted by
jmax , is necessarily given by j1 +j2 , the sum of the two maximum values of m1
and m2 . Notice that for j = j1 +j2 one can realise all values of m ranging from
−j to j. Indeed the case m = −j corresponds to m1 = −j1 and m2 = −j2 .
We can then lower j by one unit obtaining states with j = j1 + j2 − 1 and so
on. The minimum value jmin can be found imposing that the total-j basis has
a number of eigenvectors equal to the number of eigenvectors in the product
basis. As we have seen, this is given by (2j1 + 1)(2j2 + 1). Considering
moreover that for each allowed value of j, there are 2j + 1 eigenvectors
corresponding to a different value of m in the range −j ≤ m ≤ j, we can
impose:
jmax
X
(2j + 1) = (2j1 + 1) (2j2 + 1) . (4.103)
jmin
and since jmax = j1 + j2 , with easy algebraic steps, one finds jmin = |j2 − j1 |.
Therefore, summarising, we have just proved the important result
|j2 − j1 | ≤ j ≤ j1 + j2 . (4.105)
6
This can be understood considering that the product basis eigenstates | m1 , m2 i are
also eigenstates of Jˆz with eigenvalue m1 + m2 . Therefore, the subspace for a fixed m,
must also correspond to the subspace with basis given by all eigenstates | m1 , m2 i with
m1 + m2 = m. The simplest case is when m1 = j1 and m2 = j2 since in this case there
is only the direct product basis eigenstate | j1 , j2 i and, therefore, the subspace is one-
dimensional. For this reason, one necessarily has | j, j i = | j1 , j2 i. The same it is true
when m1 = −j1 and m2 = −j2 . In this case one has | j, −j i = | − j1 , −j2 i. We will see
this algebraically when we discuss the Clebsch-Gordan coefficients.
122 CHAPTER 4. ANGULAR MOMENTUM
If we now take the inner product with h j1 m1 , j2 m2 | and act with Jˆz on the ket and with
(1) (2)
−Jˆz − Jˆz on the bra, we immediately obtain
(m − m1 − m2 ) h j1 m1 , j2 m2 | jm i = 0 ,
We have now to recall the general result Eq. (4.19) for Jˆ−
Jˆ− | j, m i = [j(j + 1) − m(m − 1)]1/2 ~ | j, m − 1 i , (4.114)
that, applied to our case, gives:
√
Jˆ− | 1, 1 i = 2 ~ | 1, 0 i . (4.115)
At the same time the right-hand side of Eq. (4.113), using (4.114) with
j = 1/2 and m = 1/2, gives
(Jˆ(1) + Jˆ(2) ) | + + i = Jˆ(1) | + + i + Jˆ(2) | + + i (4.116)
p p
= ~ ( 3/4 + 1/4 | − + i + 3/4 + 1/4 | + − i)
= ~ (| − + i + | + − i) , (4.117)
obtaining
| − +i + | + −i
| 1, 0 i = √ , (4.118)
2
124 CHAPTER 4. ANGULAR MOMENTUM
In the product space V1⊗2 , the direct product basis vectors will be represented
by four components column vectors:
1 0 0 0
0 1 0 0
| ++ i → 0 , | +− i → 0 , | −+ i → 1 , | −− i → 0 .
0 0 0 1
(4.124)
If we now consider the representations of triplet states and of the singlet, the
total-j basis vectors, we have:
1 0 0 0
0 1 1
1 1
0 .
0 , | 1, 0 i → √2 1
| 1, 1 i → | 0, 0 i → √ | −1 i →
, , 1,
2 −1 0
0 0 0 1
(4.125)
It is clear that the components of the two states | 1, 0 i and | 0, 0 i cannot be
written as the product of the components of two one particle states, simply
because the first and fourth component vanish and this implies for the first
(1) (2)
component that either a1 = 0 or a1 = 0 and similarly for the second
(1) (2)
component, either a2 = 0 or a2 = 0. However, then this necessarily
implies that also the second and third component of the two particle states
should vanish and one would get the null vectors. This confirms that these
two states are entangled states and cannot be written as the product of two
one-particle states.
126 CHAPTER 4. ANGULAR MOMENTUM
Bibliography
127
128 BIBLIOGRAPHY
Chapter 5
Relativistic Quantum
Mechanics
129
130 CHAPTER 5. RELATIVISTIC QUANTUM MECHANICS
If we now project on the x̂-basis, we obtain an equation for the wave function
ψ(x, t) that is clearly relativistically non invariant :
∂ ~2 2
i~ ψ(x, t) = − ∇ ψ(x, t) . (5.4)
∂t 2m
Indeed one can see that time and position are not equally treated, as one
would expect in the case of a covariant equation. Of course this is not a
surprise considering that we started from an Hamiltonian operator corre-
sponding to a non-relativistic Hamiltonian.
There are two different solutions to the problem of getting a relativistic
invariant equation: one leads to the Klein-Gordon equation and one leads to
the Dirac equation. Here we just focus on the problem of getting a relativistic
invariant equation for the wave function in position basis. We will then
comment on the fact that both equations suffer other problems that can be
solved only definitively giving up the idea of a wave function formulation of
QM, simply because it is intrinsically at odd with special relativity and it
can only provide an approximation valid if certain conditions hold.
∂
q
i~ | ψ(t) i = c2 p̂2 + m2 c4 | ψ(t) i . (5.6)
∂t
However, this equation is not yet relativistically invariant, since the left-hand
side is linear in time derivative while on the right-hand side we have a square
root. When this is projected in position basis one would obtain a square root
of a second order space derivative plus a constant that is equivalent to have
a series of terms containing space derivatives to all orders. This means that
5.2. THE DIRAC EQUATION 131
the equation would not only be covariant, but it would not even be local, i.e.,
the value of the wave function in one point would depend on the value of the
wave function in an arbitrarily far point: this clearly would violate causality.
In order to circumvent this problem, the solution seems to be to get rid of
the square root and to this extent there are two ways to proceed. The first one
is simply to apply the time derivative on both sides and since this commutes
with the Hamiltonian operator of a free particle that is time-independent,
this is equivalent to square the operators on both sides in Eq. (5.1) obtaining
:
∂2
−~2 2 | ψ(t) i = c2 p̂2 + m2 c4 | ψ(t) i .
(5.7)
∂t
If we now project on the position basis, divide both sides by (~c)2 , bring the
left-hand side on the right-hand side and change the overall sign, we obtain
the Klein-Gordon equation:
" 2 2 #
2
1 ∂ mc
2 2
− ∇2 + ψ(x, t) = 0 . (5.8)
c ∂t ~c
Using natural units (c = ~ = 1) and introducing the D’Alambertian (or box)
operator ≡ ∂ 2 /∂t2 − ∇2 , it can be more compactly rewritten as
( + m2 ) ψ(x, t) = 0 . (5.9)
Here notice that the wave function is a scalar, i.e. a one-component wave
function. Therefore, the Klein-Gordon equation cannot describe spin 1/2
particles such as electrons, simply because, as we have seen, these have to be
described by spinors, i.e., multicomponent wave functions. There is, however,
a second solution to the problem of getting a relativistic invariant equation
for free particles and this is suitable to describe spin 1/2 particles such as
electrons.
where the four quantities α̂ ≡ (α̂1 , α̂2 , α̂3 ) and β̂ should be meant as operators
acting in some spinorial space with a dimension yet to be determined. A
priori one could think this can be just one, then the α̂i ’s and β̂ would be just
scalars and one would go back to the case of a scalar equation. However, as
we will see in a moment, this will be not the case. We now impose that when
we square both sides of (5.10), the left-hand side still respect the classical
(relativistic) analogue equation:
In this way one has to impose that α̂ and β̂ have to be such to satisfy the
condition
p̂2 + m2 = (α̂ · p̂ + β̂m)2 . (5.12)
This condition has no solution if α̂ and β̂ were just scalars. They necessarily
have to be operators satisfying the following anti-commutation conditions:
In this Chapter1 we discuss four topics that all contributed, and still con-
tribute today, to shape what can be called the second quantum revolution.2
We start first discussing Bell inequalities applied to the entangled singlet
state we obtained in the last chapter in the case of two spin 1/2 particles.
We will see how these provided the theoretical tool to definitely establish QM
description as a complete theory without the need of introducing so-called
hidden variables. We discuss then how the study of entangled states has
inspired important applications. We first discuss quantum cryptography and
then quantum computing. Finally, we discuss the important idea of quantum
decoherence as a way to understand why macroscopic systems do not exhibit
quantum behaviour such as quantum superposition and this will provide a
solution to the famous Schrödinger’s cat paradox.
135
136 CHAPTER 6. THE SECOND QUANTUM REVOLUTION
until 1964, when J.S. Bell derived an inequality (now bearing his name)
constraining probabilities of observing certain spin correlations that allows
to test whether QM should be regarded as an incomplete theory or not. In
the first case one should introduce so-called hidden variables able to reconcile
the experimental findings with Einstein’s locality principle and such that the
physical reality of the two components of the system is maintained even in
entangled states. With the advent of laser technology, experiments testing
the inequality could gradually be performed with increasing precision and
accuracy. All experimental results have unambiguously ruled out alternative
models of QM, respecting locality and physical reality hypotheses.
where we recall that the first sign refers to particle 1 and the second sign
to particle 2. The two product eigenstates | + − i and | − + i would have
a clear classical interpretation, since each particle would have a well defined
spin. However, their superposition is now telling us that when we describe
the system as a whole we need to abandon the idea that each particle has
a definite spin, we simply cannot know which particle has spin up and spin
down in such a state unless we make a measurement. This would have as an
effect that the singlet state would collapse in one of the two product states,
either | + − i or | − + i, with equal probability. In this case, if one measures
the first particle in the spin-up state, then automatically any subsequent
measurement of the state of the second particle would necessarily give spin-
down as outcome and vice versa. If the system is in one of the two product
states, the two particles gain their distinguishability. However, in the singlet
state they would be entangled together and it makes sense just to talk of
state of the whole system. For example, the two electrons in the Helium
atom can be in the singlet state and be indistinguishable. At first sight this
6.1. BELL INEQUALITIES 137
situation seems easy to accept and that does not lead to any real paradox or
inconsistency.
However, the existence of such entangled states have given risen to a very
intense debate that ultimately stimulated the developments of new quantum
technologies based on their properties. The point is that the entanglement
and the consequent correlation in the measurement of the two spins persists
even if the two particles are produced by the decay of a parent particle in the
singlet state, flying apart and becoming well separated ceasing to interact.
Let us indeed now suppose that Alice can either measure either Sz or Sx
of the first particle, while Bob always measures Sx of the second particle. Let
us remind that from Eq. (4.86) one has that the Ŝx one-particle eigenstates
138 CHAPTER 6. THE SECOND QUANTUM REVOLUTION
• Particle 1 with (Sz , Sx ) = (~/2) (+1, +1) and particle 2 with (Sz , Sx ) =
(~/2) (−1, −1);
• Particle 1 with (Sz , Sx ) = (~/2) (−1, +1) and particle 2 with (Sz , Sx ) =
(~/2) (+1, −1);
• Particle 1 with (Sz , Sx ) = (~/2) (−1, −1) and particle 2 with (Sz , Sx ) =
(~/2) (+1, +1) .
One can notice that for all four states, the measurement of one of the two
spins in particle two would give as outcome an opposite value, as in the
quantum-mechanical case, since this is enforced by angular momentum con-
servation. In this model, if Bob decides to measure Sz of particle 2, he will
get a result that is independent of whether Alice decides to measure Sx or
Sz . In this way the locality principle is incorporated in the model: Alice
and Bob results are pre-determined, independently of what the other has
measured or not measured, prior to her/his measurement. The model also
respects physical reality: each particle has an intrinsic well defined value
of one of the two spins independently whether a measurement is performed
or not and independently of the spin of the other particle. However, at the
same time, the statistical results of QM are reproduced by this model, simply
since Alice and Bob perform many measurements and there is a well deviced
statistical distribution, in this case a homogeneous one, of how the different
four configurations are produced at the source. Notice here we do have an
ensemble but it is not made of many copies of the same system, rather it is
an ensemble containing copies of four different two-particle systems that are
produced, one after the other, at the source.
This model seems really to establish an intrinsic ambiguity in the inter-
pretation of the results and it seems to confirm that it is indeed possible
to conceive a model respecting locality and physical reality reproducing the
same QM spin correlation measurement results. However, if one considers
more complicated situations, Wigner’s model predict different outcomes from
QM and one can perform a crucial test to unambiguously understand which
model is correct. As we have seen, one can consider spin projection along
142 CHAPTER 6. THE SECOND QUANTUM REVOLUTION
any arbitrary direction (see Eq. (4.88)). Let us consider then three arbi-
trary directions n̂1 , n̂2 and n̂3 . In QM we can perform measurements along
these three directions, in Wigner’s model we assign to each particle a value
Sni = ~/2 or −~/2 for each direction n̂i . Clearly we have this time eight
possible different combinations of values (Sn1 , Sn2 , Sn3 ) let us say for particle
1. Again we can consider a singlet state where particle 2 will have values
(−Sn1 , −Sn2 , −Sn3 ), i.e., opposite to those of particle 1. Again, at the pro-
duction the two-particle systems are produced in the singlet state in one of
these eight possible combinations with a certain distribution of probabilities.
Despite the arbitrariness in these distributions it is possible to derive the
following Bell’s inequality for the probabilities P (Sni , Snj ) with i, j = 1, 2, 3
and i 6= j that Alice measures Sni and Bob measures Snj :
~ ~ ~ ~ ~ ~
P Sn1 = , Sn2 = ≤ P Sn1 = , Sn3 = + P Sn3 = , Sn2 = .
2 2 2 2 2 2
(6.5)
This is the prediction from Wigner’s model respecting the locality and phys-
ical reality hypotheses.
On the other hand in QM it is possible to derive the following equalities:
~ ~ 1 2 θij
P Sni = , Snj = = sin , (6.6)
2 2 2 2
where θij is the angle between n̂i and n̂j , i.e., such that n̂i · n̂j = cos θij . If
we now plug Eq. (6.6) in Bell’s inequality (6.5), the latter specialises into:
2 θ12 2 θ13 2 θ32
sin ≤ sin + sin . (6.7)
2 2 2
However, it is simple to see that this inequality cannot be satisfied for all
possible directions n̂1 , n̂2 and n̂3 , implying that there are choices of the three
angles θ12 , θ13 and θ32 for which the inequality is clearly not holding. For
example, taking θ12 = 2 θ and θ13 = θ32 = θ, the inequality is violated
for 0 < θ < π/2.5 Therefore, the QM predictions are not compatible with
Bell’s inequality. This shows that there is an intrinsic diffrence between
QM and any model one can conceive where locality and physical reality
hypotheses are satisfied. This however does not mean that QM violates
5
For example, choosing θ = π/4, one obtains from Eq. (6.7) the absurd result 0.5 ≤
0.292.
6.2. QUANTUM COMPUTING 143
causality and, for example, one cannot use spin correlation measurements to
transmit information between two points faster than the speed of light.
Alice, though after the measurement of Sz knows what would Bob would
obtain as outcome if he also measures Sz , cannot pre-determine, before her
measurement, what Bob will obtain in a way to send him a signal. Bob, upon
repeated measurements of Sz , will obtain a random sequence of positive and
negative values, without even knowing whether Alice has measured or not Sz
prior to his measurements. Of course they can collect the data, reconvene,
compare each other findings and verify that there is indeed a strong corre-
lation. The same conclusion would hold even if Alice, during the series of
measurements, changes her mind and start to measure Sz after having agreed
with Bob that both would have measured Sx : there would be no way how
Bob could understand that Alice has violated their agreement, though now
the two series of measurements would become uncorrelated since the time
Alice decides to measure Sz instead of Sx . In conclusion, quantum entan-
glement, cannot be used to use a kind of super-luminal transmitter, as in
science-fiction novels.
6
The bit can be mathematically regarded as the generic element of the finite field
GF (2).
7 n
A string {ni } of n bits, with ni = 0, 1 and i = 0, . . . , n − 1, corresponds to an
Pn−1
integer number in decimal notation that is simply given by N = i=0 ni 2i . For example:
111 → 7, or 1111 → 15. You can find a converter on the web.
144 CHAPTER 6. THE SECOND QUANTUM REVOLUTION
long keys, it is currently used only for most critical applications. Standard
cryptosystems of this kind are DES, IDEA and AES.
The critical stage in the one-time pad method is the distribution of the
two copies of the keys. It involves secret channels that might be intercepted
by Eve with technologies more advanced than those of the sender and in-
tended receiver. Here is where quantum cryptography basic idea comes to
the rescue. The security of transmission rests on QM postulates. Since in
QM all measurements perturb the system in some way, it is possible to detect
Eve’s action by identifying the trace necessarily left on the transmitted key.
In the absence of such a trace one can be certain that the message has passed
without having been read by Eve. This simple idea can be summarised as:
No perturbation ⇒ No measurement ⇒ No eavesdropping
The first protocol of quantum cryptography was proposed in 1984 by
C.H. Bennett and G. Brassard, hence the name BB84. A variation of the
BB84 protocol is the EPR protocol (also called E91 protocol since it was
proposed by Artur Ekert in 1991). The idea consists in having a quantum
channel carrying the key in the form of entangled particles sent not from Al-
ice to Bob but one particle to Alice and one to Bob from a common source,
similarly to what happens in the case of the EPR paradox setup with spin-
correlation measurements, as we discussed in the previous chapter. If there
is no eavesdropping during the transmission, entangled particles arrive to
Alice and Bob who, performing a series of measurements, will find random
but perfectly correlated results. As long as they have not done the mea-
surements, the results are not predictable and the key simply does not exist.
Comparing their measurements, they can test whether the data violate Bell’s
inequality, as entangled states do, or whether during the transmission Eve
has intercepted the secrete key making the entangled state to collapse in
product states.
The time evolution of the state vector is described by the Schrödinger equa-
tion. An alternative description of the state of the system, introduced by
von Neumann and independently by Landau in 1927, is given by the density
operator associated to the normalised state vector | ψ i that is defined as:
ρ̂ ≡ | ψ ih ψ | . (6.14)
Notice that this is also the definition of the projector operator on the state
| ψ i given in Eq. (1.102), therefore, the density operator is the projection
operator on the state. In a given basis | i i we can of course also introduce
the associated density matrix ρ with matrix elements simply given by
states | ψk i, the density operator for such mixed state is then given by
X
ρ̂ = wk | ψk ih ψk | (6.18)
k
Notice that here the sum over k indicates an incoherent sum over the different
state vectors while indexes i and j label the basis vectors and the components
of | ψ i on these basis vectors. Some of the properties of the density matrix
valid for pure states are still valid for mixed states, specifically the density
matrix is still Hermitian (ρ̂† = ρ̂), the trace of the density matrix is still
unity Tr[ρ] = 1, its dynamics is still described by the Liouville-von Neumann
equation (6.23) and, importantly, the expectation value of an operator can
be still calculated using (6.17). However, for mixed states now ρ̂2 6= ρ̂, as it
can be easily seen again from the matrix elements in Eq. (6.19).
As an example one can build the density matrix for electrons in eigenstate
of Ŝx . In particular, one can easily see that the eigenstate with positive spin,
Sx = ~/2, represented in the Ŝz basis (see Eq. (4.86)), corresponds to a
density matrix
1/2 1/2
ρ= . (6.20)
1/2 1/2
and one can verify that ρ2 = ρ. On the other hand, in the Ŝx basis, it would
simply be
1 0
ρ= . (6.21)
0 0
This situation of course describes a beam of electrons fully polarised along
the x axis. Suppose now that the electrons interact with some walls or with a
inhomogeneous magnetic field that tends to randomise the spins. In this way
the beam will gradually become unpolarised and the density matrix would
gradually make a transition
1/2 1/2 1/2 0
ρ= −→ , (6.22)
1/2 1/2 0 1/2
where the asymptotic limit on the right hand side corresponds to the density
matrix for a fully unpolarised beam. Indeed it is proportional to the identity
6.4. QUANTUM DECOHERENCE 151
and it is invariant under rotations, meaning that it would look the same
in any basis. This means that independently on which spin component is
measured, one would always find that half of the electrons have spin up and
half have spin down, so that the polarisation, the average spin, vanishes.
The behaviour in Eq. (6.22) is called decoherence and it is a very general
process happening when a microscopic system interacts with a macroscopic
one. The transition can be described by the following Lindblad equation [10]
∂ρ 0 1
i~ = [H, ρ] − D , (6.23)
∂t 1 0
where D is the decoherence rate and τdec is the time of relaxation to decoher-
ence of the system. The off-diagonal terms are exponentially damped by the
decoherence term and, therefore, asymptotically the density matrix becomes
diagonal. For macroscopic systems τdec is tiny and this is the reason why
macroscopic system are instantaneously projected on some eigenstate and
are never observed in a superposition state.
This decoherence effect solves the famous Schrödinger’s cat paradox: “A
cat is placed in a steel chamber, together with the following hellish contrap-
tion . . . In a Geiger counter there is a tiny amount of radioactive substance,
so tiny that maybe within an hour one of the atoms decays, but equally
probably none of them decays. If one decays then the counter triggers and
via a relay activates a little hammer which breaks a container of cyanide. If
one has left this entire system for an hour, then one would say the cat is
living if no atom had decayed. The first decay would have poisoned it. The
wave function of the entire system would express this by containing equal
parts of the living and dead cat.”
Therefore, after an hour the state would be described by a state vector
of the schematic form
1
| ψ i = √ (| ψalive i + | ψdead i) . (6.24)
2
The cat would then be in a superposition of dead and alive states. Opening
the box would then correspond to a measurement process and there would
be 50-50 chance to find it dead or alive! Obviously this was a burlesque way
to put the question why macroscopic systems are never observed in a super-
position states. We know the answer now: because the environment itself,
independently whether there is an observer or not, continuously makes the
state of a macroscopic system collapsing in such a fast way that it can actu-
ally never be observed in a superposition state and repeated measurements
152 CHAPTER 6. THE SECOND QUANTUM REVOLUTION
of the same state will always give the same outcome: in this way a classical
behaviour is recovered. Schrödinger’s cat, by the way, still lived a long life!
Bibliography
[4] For a review on the experiments that have been performed to test Bell’s
inequalities see A. Aspect, Bell’s inequality test: more ideal than ever,
Nature 398, 189–190 (1999), https://doi.org/10.1038/18296.
153
154 BIBLIOGRAPHY