Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

1 Quantum States

Quantum Mechanics describes the behaviour of matter and light at the atomic
scale, where physical systems behave very differently from what we experience in
everyday’s life – the laws of physics of the quantum world are different from the
ones we have learned in Classical Mechanics. Despite this “unusual” behaviour,
the principles of scientific inquiry remain unchanged: the only way we can access
natural phenomena is through experiment; therefore our task in these first chapters
is to develop the tools that allow us to compute predictions for the outcome of
experiments starting from the postulates of the theory. The new theory can then
be tested by comparing theoretical predictions to experimental results. Even in the
quantum world, computing and testing remain the workhorses of physics.
Because Quantum Mechanics is so different from our intuitive, “classical” de-
scription, the procedure we adopt for describing the quantum world needs to be
defined carefully; the physical description of quantum phenomena must adhere very
strictly to this procedure. In the first part of this book we will focus on presenting
the basic postulates, and in developing the mathematical tools needed for studying
elementary systems. In time we want to be able to relate the mathematical con-
structs to actual physical processes. As we venture on this journey, we can only
develop some intuition by practice; solving problems is an essential component for
understanding Quantum Mechanics.
Experimental results at the beginning of the 20th century first highlighted that
the behaviour of physical systems at atomic scales are inconsistent with Classical
Mechanics. As discussed in the previous chapter, the combined effort of many em-
inent physicists led to a consistent picture of the new dynamics. Here we will not
follow the historical route; unsurprisingly the latter was rather tortuous. Instead
we will start by stating the new principles, and then work towards deriving their
consequences.
Despite its unintuitive aspects, Quantum Mechanics describes very concrete fea-
tures of the world as we know it, e.g. the spectrum of the hydrogen atom. By now,
many of its predictions have been tested to great accuracy, and have always been
found in agreement with experiments. This book’s ambition is to set the founda-
tions for developing a description of natural phenomena as observed in nature, i.e.
to connect the abstract to the experiment. At every stage we will try to provide
examples of physical systems described by the techniques that are being discussed.
i
ii Quantum States

It is useful to keep in mind the following disclaimer in Feynman’s lectures 1 :


“In this subject we have, of course, the difficulty that the quantum mechanical
behavior of things is quite strange. Nobody has an everyday experience to lean on to
get a rough, intuitive idea of what will happen. So there are two ways of presenting
the subject: We could either describe what can happen in a rather rough physical way,
telling you more or less what happens without giving the precise laws of everything;
or we could, on the other hand, give the precise laws in their abstract form. But, then
because of the abstractions, you wouldn’t know what they were all about, physically.
The latter method is unsatisfactory because it is completely abstract, and the first
way leaves an uncomfortable feeling because one doesn’t know exactly what is true
and what is false. [...] Here, we will try to find a happy medium between the two
extremes”.
Any first course in Quantum Mechanics suffers from the same problem, and this
textbook makes no exception. We chose to start with an abstract definition of the
laws, making use when possible of abstract mathematical notation. The explicit
examples should allow the reader to appreciate the power of the mathematical
tools, while making the mathematical notation more meaningful.
In order to start our journey into Quantum Mechanics, we need to introduce two
main concepts:
1. the state of the system, and later in this book the dynamical laws that determine
its time evolution;
2. the observables, i.e. the possible outcomes of experiments, which allow us to
probe the state of the system.
This chapter is devoted to the introduction of the mathematical structure needed
to characterise the state of a quantum system, while the study of observables is
deferred to the following chapter in the book. Here we will see that physical states
are associated to vectors in complex vector spaces, and we will define the main
concepts that are going to be used to deal with physicals states. These are the
foundations of the theory and will be used throughout the book. Different phys-
ical systems will have states belonging to different vector spaces, but the general
mathematical framework will remain the same for all quantum system. The tools
that we introduce in this chapter will then be used for all subsequent discussions.
Examples and exercises should help you in building the skills needed to manipulate
vectors in cases that are relevant for Quantum Mechanics.

1.1 States of a Quantum System

In Classical Mechanics, the state of a point-like system is described by the position


and the momentum of its components. For an elementary particle in our usual
1 R.P. Feynman, R.B. Leighton, M. Sands, The Feynman lectures on physics - Quantum Me-
chanics, Addison-Wesley, 1965.
iii States of a Quantum System

three-dimensional space, the state of the system is defined by a six-dimensional


real vector:

x, p . (1.1)

Underlined letters indicate three-dimensional vectors, x and p being respectively the


position and the momentum of the particle. The knowledge of these two vectors at
any moment in time allows us to compute all other properties of the system, like
e.g. its energy or its angular momentum. Moreover, once the dynamical laws are
given, the state of the system at any other time can also be computed by solving
the equations of motions.
We shall now introduce the postulates that allow us to describe the state of
a quantum system. It is useful here to focus on the peculiarities of the quantum
description. We will try to highlight these peculiarities as we encounter them and
to reinforce our understanding through examples and exercises. We should always
be able to connect the abstract concepts that are going to be introduced to physics,
i.e. use the mathematical tools to compute theoretical predictions for the outcome
of experiments. This process needs to be carefully defined, we will need to specify
exactly what quantities can be computed and how these calculations are performed.
The first postulate of Quantum Mechanics defines the states of a quantum sys-
tem.

Postulate 1 Quantum states are identified by vectors in complex vector spaces.

It is important to emphasize that we will deal with complex vector spaces, instead
of the real vector spaces that are more familiar from Classical Mechanics. This
entails some differences that will be discussed in this chapter. A first comment
arises by comparing the state vector in Classical and Quantum Mechanics. As we
discussed in the example above, the state of a classical pointlike system is described
by its position and its momentum, a six-dimensional real vector. In this case the
components of this six-dimensional vector are also observables: the position and
the momentum of the particle can be measured in an experiment. In Quantum
Mechanics the picture is more subtle. Experiments do not measure complex vectors
and we will have to put in some more work to define what can be predicted about
the outcome of experiments.
For now, let us begin by introducing the concepts from linear algebra that are
needed to work with the elements of complex vector spaces. This is standard ma-
terial to be found in linear algebra courses, which we summarise in order to have a
consistent set of definitions and notations. The chapter is dotted with examples and
simple exercises that will hopefully help the reader to check that they understand
and are able to apply the concepts that are being introduced. Their connection
to physics will be the subject of following chapters. As we introduce physical con-
cepts later in this book, we will continuously refer to the material presented in this
chapter.
iv Quantum States

1.1.1 Kets

Following Dirac’s notation, vectors will be denoted by kets, so e.g. the elements of
a vector space H are written as
|vi ∈ H . (1.2)

More generally we will use the notation | . . .i, where . . . can be any combination
of symbols used to identify the vector. Sometimes we will insert just one letter,
e.g. |ψi, some other times we will use one or more indices, e.g. |ni, or |n, mi, with
n, m integers. Finally we will encounter instances where the kets are identified by a
continuous index, e.g. |ξi with ξ ∈ R. We will use the words state, vector, and ket
interchangeably.
The defining property of a vector space is that linear combinations of vectors also
belong to the vector space:

∀|ui, |vi ∈ H, ∀α, β ∈ C, α|ui + β|vi ∈ H . (1.3)

The fact that H is a complex vector space is reflected in the fact that the coefficients
α and β are complex numbers. It is straightforward to generalise this result to the
linear combination of a set of vectors indexed by some integer n:
X
∀ {|un i} ⊂ H, ∀ {αn } ⊂ C, αn |un i ∈ H . (1.4)
n

In Eq. (1.4) we have deliberately not specified the range of values taken by the
index n. Depending on the system under study we may deal with either finite or
infinite sums. We will come back to this issue later in the chapter.

Mathematical aside: properties of vector spaces


Here are some useful rules for manipulating vectors. In the following formulae, kets
indicate generic elements of H and coefficients are generic complex numbers. Some
of these properties may look familiar to readers who have some experience with
linear algebra.

• There exists a ‘+’ operation between vectors:

∀|ψi, |φi ∈ H, |ψi + |φi ∈ H . (1.5)

We have already implicitly used this property in Eq. (1.3).


• The sum of vectors is associative and commutative:

|ψi + (|φi + |ξi) = (|ψi + |φi) + |ξi , (1.6)


|ψi + |φi = |φi + |ψi . (1.7)

Therefore the brackets in the first line are superfluous, and one can write simply

|ψi + |φi + |ξi . (1.8)


v States of a Quantum System

• Multiplication of a vector by a complex number yields another vector:

∀|ψi ∈ H, ∀α ∈ C, α|ψi ∈ H . (1.9)

We will typically write the complex coefficient to the left of the vector, but
this is just a convention:

α|ψi = |ψiα . (1.10)

• The multiplication by complex numbers is associative and distributive:

α (β|ψi) = (αβ) |ψi , (1.11)


α (|ψi + |φi) = α|ψi + α|φi , (1.12)
(α + β) |ψi = α|ψi + β|ψi . (1.13)

Let us examine in more detail the operations that are described in the first of
these equations. Note that, on the left-hand side of Eq. (1.11), we first perform
the multiplication of a vector, |ψi, by a complex number β. The result of this
operation is a vector, which is then multiplied by the complex number α. The
final result on the left-hand side is therefore a vector. On the right-hand side,
we first multiply the two complex numbers α and β. The result of this operation
is another complex number, (αβ), which multiplies the vector |ψi. The result
on the right-hand side is also a vector, as one would expect if the equality
has to hold. As a consequence of Eq. (1.11) the expression αβ|ψi can be used
without ambiguities. It is useful to perform this kind of ’grammatical’ analysis
of the equations, until the reader feels completely familiar with the operations
performed with vectors in complex vector spaces.
• Existence of a null vector |0i such that:

|ψi + |0i = |ψi , (1.14)


0|ψi = |0i , (1.15)
α|0i = |0i . (1.16)

• For each vector |ψi, there exists an inverse vector with respect to the ‘+’ opera-
tion, denoted −|ψi’:

|ψi + (−|ψi) = |0i . (1.17)

Vector spaces are ubiquitous in physics. The most familiar examples being the
real vector spaces used in Classical Mechanics. Despite the abstract definitions
above, vector spaces encode familiar properties. It is worthwhile to discuss a simple
example of a complex vector space, which allows us to illustrate the main concepts.
vi Quantum States

Example 1.1 Consider the set of complex column vectors of size 2, i.e. the set of
column vectors that can be written as
 
v1
|vi ≡ , (1.18)
v2

where v1 and v2 are complex numbers. This space is called C2 . If we define the sum
of vectors and the multiplication by a complex scalar λ respectively as
     
v1 w1 v + w1
+ = 1 , (1.19)
v2 w2 v2 + w 2
   
v1 λv1
λ = , (1.20)
v2 λv2

then C2 is a vector space. It is important to notice that the ’+’ sign on the left-
hand side of Eq. (1.19) defines the sum of two column vectors. The ’+’ sign on the
right-hand side indicates that the components of each column vector are summed,
and therefore is simply the sum of complex numbers that should be familiar to the
reader. The reader can check that the elements of C2 satisfy all the properties listed
above.

Superposition Principle
The vector space structure automatically implements the superposition prin-
ciple, which states that any linear combination of quantum states with complex
coefficients defines a new, valid quantum state of the system. This property will
recur multiple times in the rest of the book.
Alternatively, one could state the superposition principle as the first fundamental
postulate and deduce that the states of a quantum system need to be elements of
a vector space.

1.1.2 Scalar product

We will always be concerned with vector spaces on which a scalar product is


defined. The scalar product, which is sometimes also called internal product, is
a function that takes two vectors, |ui and |vi, and returns a complex number,
which we will denote by hu|vi. We can rephrase this sentence using a mathematical
notation and define the scalar product as:

H×H→C
|ui, |vi 7→ hu|vi .

The notation above simply indicates that the scalar product is a function that takes
two vectors and returns a complex number. It satisfies the following properties:
vii States of a Quantum System

1. Complex conjugation:

∀|ui, |vi ∈ H, hu|vi = hv|ui∗ . (1.21)

2. Linearity:
h i
∀|ui, |vi, |wi ∈ H, and ∀α, β ∈ C, hu| α|vi + β|wi = αhu|vi + βhu|wi . (1.22)

Two comments are in order here. First, the scalar product is not symmetric with re-
spect to the permutation of its arguments, hu|vi =6 hv|ui. Second, note that because
of Eq. (1.21) above, the scalar product is antilinear in its first argument, viz.
h i
αhu| + βhv| |wi = α∗ hu|wi + β ∗ hv|wi . (1.23)

Vector spaces equipped with a scalar product are called Hilbert spaces.

Example 1.2 Given two vectors


   
v1 w1
, ∈ C2 , (1.24)
v2 w2
the scalar product can be defined as

hv|wi = v1∗ w1 + v2∗ w2 . (1.25)

Check that this definition satisfies the properties above.

1.1.3 Bras

A linear functional Φ is a function defined on H that associates to each vector a


complex number. Using a mathematical notation, we can write:

Φ:H→C
|vi 7→ Φ (|vi) = z .

Linearity means that Φ satisfies

Φ (α|vi + β|wi) = αΦ (|vi) + βΦ (|wi) , (1.26)

for arbitrary vectors |ui, |vi and complex coefficients α, β.


For a given vector |ui, the scalar product associates a complex number to any
vector |vi through the correspondence

|vi 7→ z = Φ|ui (|vi) = hu|vi , (1.27)

and therefore defines a bona fide functional Φ|ui acting on H. The linearity of the
functional in this case is guaranteed by Eq. (1.22). It can be shown that all bounded
functionals acting on H can be defined as in Eq. (1.27) - the latter result is known
viii Quantum States

as Riesz theorem, whose proof is beyond the scope of these lectures. The one-to-
one correspondence between functionals and vectors led Dirac to introduce the bra
notation, hu|, to denote the functional defined in Eq. (1.27). The notation hu|vi can
also be seen as the application of the functional hu| to the vector |vi. The space of
functionals is called the dual of the vector space H, and is denoted by H∗ .

Example 1.3 We can look once again at the C2 vector space, and understand in
detail the action of a functional. In order to have an explicit example, let us choose
the vector
 √ 
i/ √2
|ui = ∈ C2 , (1.28)
−1/ 2
and consider the functional hu| defined in Eq. (1.27). Acting with hu| on
 
1
|v1 i = (1.29)
0
yields

hu|v1 i = −i/ 2 ∈ C . (1.30)
Take a moment to make sure you understand that the functional acts on a vector
and returns a scalar. Check that you can reproduce the result in the above equation.
What is the origin of the minus sign on the right-hand side?
Acting with hu| on a generic vector
 
v
|vi = 1 , v1 , v2 ∈ C (1.31)
v2
we obtain
i 1
hu|vi = − √ v1 − √ v2 . (1.32)
2 2
Note that the result of the action hu| on a vector returns a complex number, as
expected for a functional.

1.1.4 Norm

As usual for spaces with a scalar product, the norm of a vector |vi will be given
by the scalar product of the given vector with itself, ||v||2 = hv|vi. For reasons
that we will discuss later, quantum states are associated to vectors with unit norm,
hv|vi = 1. Vectors with non-unit norm can be normalised:
1
|vi 7→ p |vi . (1.33)
hv|vi
You should practice the use of Eq. (1.33), as it will be used numerous times in the
rest of the book, and in real life applications.
ix States of a Quantum System

Example 1.4 For all vectors


 
v1
∈ C2 , (1.34)
v2
the norm obtained using the scalar product defined above is

||v||2 = v1∗ v1 + v2∗ v2 . (1.35)

You should convince yourself that the norm is a positive real number. It is in-
structive to compute explicitly the norm of a few vectors in C2 .

Exercise 1.1.1 The vector


 
1
|vi = ∈ C2 (1.36)
2i
has norm squared hv|vi = 5. Normalise the vector so that it has unit norm.

We will encounter vectors of infinite norm, which require us to extend the prop-
erties of the scalar product in order to be able to deal with them. In particular,
if |ξi is a vector of infinite norm, we shall assume that its scalar product with a
vector |vi of finite norm is finite.
The norm satisfies the following triangular inequalities:

||u + v|| ≤ ||u|| + ||v|| , (1.37)


||u|| − ||v|| ≤ ||u + v|| . (1.38)

Exercise 1.1.2 Consider the vectors


   
1 1
|ui = , |vi = ∈ C2 (1.39)
0 2i
and check that the triangular inequalities are satisfied. Make sure you can
add and subtract vectors, and understand how the norm is computed in this
simple case.

Rays and Phase Ambiguity


A more sophisticated wording to describe the quantum-mechanical states is to say
that they are associated to rays in the vector space. Note that the condition hv|vi =
1 leaves the vector undefined up to a phase factor eiϕ . We have the following
important result: vectors that differ by a global phase factor describe exactly the
same quantum state. As we will see in the following chapters, physical predictions
are always independent of the phase of the ket.
x Quantum States

1.1.5 Orthogonality

Two vectors |ui and |vi are orthogonal if hu|vi = 0. Two subspaces H1 , H2 in H
are orthogonal subspaces if

∀|ui ∈ H1 , |vi ∈ H2 , hu|vi = 0 . (1.40)

Given a subspace H1 in H, the set of vectors orthogonal to H1 is also a subspace,


called the complementary subspace to H1 . We shall denote the complementary
subspace by H1,⊥ . For each vector |vi ∈ H there is a unique decomposition:

|vi = |v1 i + |v⊥ i , |v1 i ∈ H1 , |v⊥ i ∈ H1,⊥ . (1.41)

The vector |v1 i is called the projection of |vi in H1 . A pictorial representation of


the projection of a vector in different subspaces is show in Fig. 1.1.

H1
|vi

|v1 i
H1,⊥

|v⊥ i

t
Fig. 1.1 A vector space H1 , and its complementary subspace H1,⊥ . The vector |vi is decomposed
into the sum of |v1 i, its projection onto H1 , and |v⊥ i, its projection onto the complementary
subspace H1,⊥ .

1.1.6 Operators

Following the formalism that we have set up so far, we expect that a change in the
state of a system is represented by a change in the state vector. Therefore we need
to introduce a mathematical tool that transforms the vectors in the Hilbert space.
Such a tool is called an operator.
More precisely, an operator is a function that acts on a vector and returns a
vector:

Ô : H → H
|vi 7→ |wi = Ô|vi . (1.42)

Both |vi and |wi are vectors in H. Note that the result of acting on a state vector
with an operator produces a new state vector. Also note the difference between an
operator and a functional. Both act on vectors, i.e. on elements of H. The action
xi States of a Quantum System

of an operator on a vector returns a vector, while the action of a functional on a


vector returns a complex number.
A quick comment about conventions is in order. We will denote operators with
“hats”, like e.g. Ô in the example above. Unless otherwise specified, we will always
use the convention that operators act from the left on to kets.
In Quantum Mechanics we will deal with linear operators. A linear operator
satisfies:
 
Ô c1 |ψ1 i + c2 |ψ2 i = c1 Ô|ψ1 i + c2 Ô|ψ2 i , (1.43)

where c1 and c2 are complex numbers.

Commutators
The product of two operators is defined as:
 
Ô1 Ô2 |ψi = Ô1 Ô2 |ψi . (1.44)

Note that the order in which the operators are applied to the state is important!
The commutator of two operators is:
h i
Ô1 , Ô2 = Ô1 Ô2 − Ô2 Ô1 . (1.45)

In general the commutator does NOT vanish; it defines an operator acting on


quantum states:
h i
Ô1 , Ô2 |ψi = Ô1 Ô2 |ψi − Ô2 Ô1 |ψi . (1.46)

Functions of Operators
The product of operators can be extended to the case of n factors in a straightfor-
ward manner:
   
Ôn |ψi = Ô . . . Ô |ψi . . . . (1.47)
| {z }
n factors

Having defined monomials of arbitrary order, we can define a function of a generic


operator using the function’s Taylor expansion. If f (x) is a function such that
X
f (x) = cn xn , (1.48)
n

then we define the action of the function on the operator:


X
f (Ô) = cn Ôn , (1.49)
n

where the coefficients cn are the same coefficients that appear in Eq. (1.48). As
expected, Eq. (1.49) defines an operator acting on vectors in H.
xii Quantum States

Example 1.5 The exponential has a well-known Taylor expansion:


X 1 x2
ex = xn = 1 + x + + ... . (1.50)
n
n! 2!

Following the above discussion, we define the exponential of an operator as


  X 1 Oˆ2
exp Ô = Ôn = 1 + Ô + + ... . (1.51)
n
n! 2!

Note that each term in the sum is an operator, and therefore the exponential of an
operator is also an operator.

Exercise 1.1.3 Compute the first two terms in the definition of


 
log 1 + Ô . (1.52)

You may wonder about the conditions under which the Taylor expansion is
convergent, and therefore defines a bona fide operator.

1.1.7 Bases

The vectors in a given set {|ek i : k = 1, . . . , n} are called linearly independent


if
X
ck |ek i = |0i ⇐⇒ ck = 0, ∀k , (1.53)
k

i.e. none of the vectors in the set can be written as a linear combination of the
others. The dimensionality of a vector space is the largest number of linearly
independent vectors. The dimensionality of the vector spaces that we will encounter
can be either finite or infinite, depending on the physical system under study.
A basis is a set of linearly independent vectors in H such that any vector |vi
can be expanded as a linear combination of the basis vectors. If we denote B =
{|ek i : k = 1, . . .} the set of vectors in the basis, then
X
∀|vi ∈ H, ∃{vk } ⊂ C, |vi = vk |ek i , (1.54)
k

i.e. every vector |vi can be written as a linear combination of the basis vectors
with complex coefficients. Clearly the coefficients depend on the vector |vi that
is expanded. Note that the sum in Eq. (1.54) runs over a finite set of vectors for
a finite-dimensional space, while the sum becomes an infinite sum for an infinite-
dimensional space. The complex coefficients vk are the coordinates of the vector
|vi in the basis B. Note at this stage the complete analogy between the complex case,
xiii States of a Quantum System

and the more familiar case of real vector spaces. Compare for instance Eq. (1.54),
with the familiar expansion of a vector in a real three-dimensional vector space V:
∀v ∈ V, v = v1 e1 + v2 e2 + v3 e3 , (1.55)

where vi are real numbers, the usual coordinates of v in the basis e1 , e2 , e3 . We
will use Eq. (1.54) everywhere in this book.
We will often work with orthonormal bases, so that for any pair of vectors in
the basis
hek |el i = δkl . (1.56)
Under this hypothesis the coordinates vk can be readily computed,
vk = hek |vi , (1.57)
and therefore we can write Eq. (1.54) as
X X
|vi = hek |vi|ek i = |ek ihek |vi . (1.58)
k k

Remember that hek |vi is a complex number, and therefore can be written equiva-
lently to the left or to the right of the vector |ek i, as discussed above in Eq. (1.10).

Example 1.6 The vectors


   
1 0
, (1.59)
0 1
form an orthonormal basis in C2 . Every vector in C2 can be expanded as
     
v1 1 0
= v1 + v2 . (1.60)
v2 0 1

The norm of the vector |vi can be computed starting from Eq. (1.58).
||v||2 = hv|vi (1.61)
X  X 
= hek0 |hv|ek0 i hek |vi|ek i (1.62)
k0 k
X
= hek |vihv|ek0 ihek0 |ek i (1.63)
k,k0
X X 2
= vk∗ vk = |vk | . (1.64)
k k

Eq. (1.64) confirms that the norm is a positive number, which vanishes only when
|vi is the null vector. Note that the complex conjugate vk∗ = hv|ek i is necessary to
guarantee that the norm is positive, as discussed above.
The second equality in Eq. (1.58) illustrates nicely how Dirac’s notation captures
xiv Quantum States

features of linear vector spaces in a suggestive graphical manner. The object |ek ihek |
can be recognised as the projector over the basis vector |ek i. We can act with |ek ihek |
to the left of a vector |vi in order to get its projection in the direction of |ek i. It is
instructive to translate this sentence into an equation:

     
|ek ihek | |vi = |ek i hek |vi = hek |vi |ek i . (1.65)

Eq. (1.65) shows that the parentheses are unnecessary in this expression, and indeed
we will not use them unless we want to emphasize a specific interpretation of the
formula. We shall denote by

P̂k = |ek ihek | (1.66)

the projector on a state |ek i. Projectors verify the usual properties:

P̂k2 = P̂k P̂k P̂` = 0 , if k 6= ` . (1.67)

Exercise 1.1.4 Verify that you recognize the properties in Eq. (1.67) by con-
sidering the more familiar case of projectors in three-dimensional Euclidean
space.

Exercise 1.1.5 Write the relations in Eq. (1.67) using Dirac’s notation for the
projectors.

Because it is valid for any vector |vi ∈ H, Eq. (1.58) can be rewritten as

X
|ek ihek | = 1 , (1.68)
k

where 1 here denotes the identity operator acting on H. Eq. (1.68) is called a
completeness relation. The vectors of a basis are said to be a complete set in
H. In a vector space of finite dimension n, any set of n linearly independent vectors
is complete. If B1 is a subset of B spanning a subspace H1 , then the projector on
this subspace is

X
P1 = |ek ihek | . (1.69)
|ek i∈B1

Exercise 1.1.6 Show that P12 = P1 , as expected for a projection operator.


xv States of a Quantum System

Solution:
  
 X  X 
P12 = (|ek ihek |) (|e0k ihe0k |)
  0 
|ek i∈B1 |ek i∈B1
X X
= (|ek ihek |) (|ek0 ihek0 |)
|ek i∈B1 |ek0 i∈B1
X X
= |ek i (hek |ek0 i) hek0 |
|ek i∈B1 |ek0 i∈B1
X X
= |ek i δkk0 hek0 |
|ek i∈B1 |ek0 i∈B1
X
= |ek ihek |
|ek i∈B1

= P1 .
Having chosen a basis B, any vector is uniquely identified by its coordinates. A
vector |vi can be represented as a column vector made of the coordinates hek |vi:
 
he1 |vi
|vi → he2 |vi . (1.70)
 
..
.
We have used an arrow instead of an equal sign to underline the fact that the vector
is represented by its coordinates only in a given basis. When there is no ambiguity,
we may use the column vector notation to represent a state, but it is necessary to
keep in mind the underlying choice of basis. It is useful to repeat the statement once
again: a vector is identified by a column vector in a given basis. Clearly the same
vector |vi can be expanded in different bases; let us assume that we have a second
basis B 0 = {|e0k i}, the coordinates of the vector will be different in each basis:
X X
|vi = vk |ek i = vk0 |e0k i . (1.71)
k k

As it is customary when performing a change of basis in a vector space, the dif-


ferent coordinates are related to each other. The relation between different sets of
coordinates can be readily worked out, the derivation being a standard result in
linear algebra. Starting from the expression for the coordinates vk0 ,
vk0 = he0k |vi , (1.72)
and using the completeness of B
X
vk0 = he0k |el ihel |vi (1.73)
l
X
= Ukl vl , (1.74)
l

where Ukl = he0k |el i defines the matrix that relates the coordinates in the two bases.
xvi Quantum States

It is important to realise that Eq. (1.73) is nothing but the usual expression for the
change of coordinates of a vector written using bras and kets. The same relation in
terms of column vectors can be written as:
 0   
v1 U11 U12 . . . v1
v 0  U21 U22 . . . v2 
 2 =   , (1.75)
.. .. .. .. ..
. . . . .
which should be familiar from linear algebra.

Exercise 1.1.7 If H is a finite vector space of dimension n, and the basis vectors
are all orthonormal, show that Ukl in Eq. (1.74) is a unitary n × n matrix.

Finally let us consider two vectors |vi and |wi, which can be respectively ex-
panded in a given orthonormal basis as:
X
|vi = vk |ek i , (1.76)
k
X
|wi = wk |ek i , (1.77)
k

then their scalar product is given by:


X
hw|vi = wk∗ vk . (1.78)
k

The proof of this statement is left as an exercise for the reader. It is important to
note that the basis vectors need to be orthogonal for the formula to hold. Also, be-
cause we are dealing with complex vector spaces, the scalar product is not invariant
under the exchange of the bra and the ket, which is reflected in the fact that the
coordinates of the bra (wk ) in Eq. (1.78) are complex conjugated.

Operators as Matrices
We can conclude this section by discussing how the action of an operator can be
implemented at the level of the coordinates of the vector. We consider a vector |vi ∈
H and a linear operator Ô. Having chosen an orthonormal basis {|ek i : k = 1, . . .},
we can expand |vi in this basis,
X
|vi = vl |el i , (1.79)
l

where vk are the complex coordinates of |vi. When acting with Ô on |vi, we obtain
a new vector,
|v 0 i = Ô|vi , (1.80)

which can also be expanded in the same basis:


X
|v 0 i = vk0 |ek i . (1.81)
k
xvii States of a Quantum System

The coordinates of |v 0 i are computed as discussed above:


X
vk0 = hek |v 0 i = vl hek |Ô|el i . (1.82)
l

Note that in order to get the second equality, we used the fact that Ô is a linear op-
erator. Eq. (1.82) is the usual matrix vector multiplication; if we define the matrix
elements of the operator Ô between two basis vectors as
Okl = hek |Ô|el i , (1.83)
then the relation between the coordinates of |vi and |v 0 i is given by the familiar
matrix-vector multiplication:
X
vk0 = Okl vl . (1.84)
l

After having chosen a basis in H, a linear operator can be represented as a matrix.


We will use an arrow to associate a matrix to an operator, in order to emphasize
again that the matrix representation depends on the basis being used:
 
O11 O12 . . .
 .. 
Ô → 
O21 . .
 (1.85)
..
.
Note that we will use matrix elements of operators not only between basis vectors
but also between generic states. The matrix element of an operator Ô between the
states |ψi and |φi is the scalar product:
hψ|Ô|φi . (1.86)

1.1.8 Tensor Product

Let us consider two vector spaces H1 and H2 . For any pair of vectors, |v (1) i ∈ H1
and |v (2) i ∈ H2 , we define the product
|v (1) v (2) i = |v (1) i|v (2) i = |v (2) i|v (1) i , (1.87)
satisfying the following properties:
(1) (1) (1) (1)
|v (1) i = α1 |w1 i + α2 |w2 i =⇒ |v (1) v (2) i = α1 |w1 v (2) i + α2 |w2 v (2) i , (1.88)
(2) (2) (1) (2)
|v (2) i = α1 |w1 i + α2 |w2 i =⇒ |v (1) v (2) i = α1 |v (1) w1 i + α2 |v (1) w2 i . (1.89)
Under these conditions, the set of linear combinations of the vectors
n o
|v (1) v (2) i; |v (1) i ∈ H1 , |v (2) i ∈ H2

spans a vector space, called the tensor product of H1 and H2 , which we denote
as
H = H1 ⊗ H 2 . (1.90)
xviii Quantum States

n o n o
(1) (2)
Given a basis B1 = |ek i in H1 , and a basis B2 = |el i in H2 , then the set
n o
(1) (2)
B = |ek el i is a basis in H, i.e.

(1) (2)
X
∀ |ψi ∈ H, ∃ ckl ∈ C : |ψi = ckl |ek el i . (1.91)
kl

Clearly if H1 and H2 have finite dimensions d1 and d2 respectively, then H is also


finite dimensional, and

dim H = d1 × d2 . (1.92)

Knowing that H is a vector space, we can introduce all the notions discussed
above (scalar product, norm, bras, operators...).
Tensor products are useful to describe the space of states of a composite system
made of two subsystems denoted 1 and 2. If the spaces of states of the subsystems
are denoted by H1 and H2 respectively, then the space of states of the whole system
is

H = H1 ⊗ H 2 . (1.93)

Operators Acting on the Tensor Product Space


Operators can be introduced in the tensor product H as functions that transform
vectors into vectors, along the lines discussed above. It is interesting to note that
to each operator Ô(1) acting in H1 we can associate an operator acting in H. If the
action of Ô(1) is such that

Ô(1) |v (1) i = |w(1) i , (1.94)

where |v (1) i, |w(1) i ∈ H1 , then we can define the action of the operator on tensor
product kets of the form |v (1) v (2) i as follows

Ô(1) |v (1) v (2) i = |w(1) v (2) i . (1.95)

Using the fact that these are linear operators, we can extend the definition of Ô(1)
to the entire space H. Strictly speaking the extension of the operator Ô(1) to the
tensor product space H defines a new operator. Nonetheless we will use the same
symbol to denote both the original operator, acting in H1 , and its extension, acting
in H. Similarly, for an operator acting in H2 :

Ô(2) |v (2) i = |w(2) i , (1.96)

where |v (2) i, |w(2) i ∈ H2 , the extension of the operator to product states is given
by

Ô(2) |v (1) v (2) i = |v (1) w(2) i . (1.97)


xix Two-state systems

Example 1.7 Let us consider an operator Ô(1) defined by specifying its action
on the basis vectors B1 :
(1) (1)
X
Ô(1) |ek i = |e(1) (1)
m ihem |Ô
(1)
|ek i (1.98)
m
(1)
X
= |e(1)
m iOmk . (1.99)
m

Note that in the first line of the equation above we have used the completeness
relation of the basis in H1 , while in the second line we have used the definition
of the matrix element given above in Eq. (1.83). Working in the basis defined in
Eq. (1.91), we can write a generic vector in H as
(1) (2)
X
|vi = vkl |ek el i , (1.100)
kl

where vkl are the coordinates of the vector in the basis B. Using the linearity
properties of the operator yields:
(1) (2)
X
|v 0 i = Ô(1) |vi = Ô(1) vkl |ek el i (1.101)
kl
(2) (1)
X X
= vkl |e(1)
m el iOmk (1.102)
kl m
!
(1) (2)
X X
= Omk vkl |e(1)
m el i . (1.103)
ml k
0
If we denote by vml the coordinates of the vector |v 0 i in the basis B,
(2)
X
|v 0 i = 0
vml |e(1)
m el i , (1.104)
ml

we see explicitly that


(1)
X
0
vml = Omk vkl . (1.105)
k

In the algebraic manipulations above, pay attention to the use of dummy indices in
the sums! The reader should repeat this exercise for the case of the operator Ô(2)
discussed above.

1.2 Two-state systems

Before going further in our setting up the mathematical tools needed for the de-
scription of states in Quantum Mechanics, it is time to consider a few physical
examples that allow us to apply the concepts introduced so far.
xx Quantum States

A simple (yet non-trivial) example of a complex vector space is a two-dimensional


vector space H. Given a basis B = {|1i, |2i}, any state in H can be written as

|vi = v1 |1i + v2 |2i , (1.106)

where v1 , v2 ∈ C. We are now going to illustrate some of the properties discussed


in the previous section in this simple example.
State vectors need to be normalized, hence
2 2
|v1 | + |v2 | = 1 . (1.107)

Because quantum states are always defined up to a complex phase we can require
without loss of generality that v1 is real. Denoting by ϕ the relative phase between
v1 and v2 , we can parametrize the state of the system in terms of two angles:

v1 = cos ϑ/2 , v2 = eiϕ sin ϑ/2 , (1.108)

where we have introduced the angle ϑ/2 to implement the normalization condition.
The parametrization above allows us to identify the quantum states with the points
on the surface of a two-dimensional unit sphere, called the Bloch sphere.

|0i

|ψi

t
|1i

Fig. 1.2 Representation of states of a two-state system as points on the surface of the Bloch sphere.

Example 1.8 It is instructive to look at a change of basis in this simple setting.


Having chosen a basis B = {|1i, |2i}, the vector |vi in Eq. (1.106) is identified by
the two complex numbers v1 and v2 . Let us now consider a new orthonormal basis,
B 0 , whose elements are:

|10 i = V11 |1i + V12 |2i , (1.109)


0
|2 i = V21 |1i + V22 |2i , (1.110)
xxi Two-state systems

where the coefficients Vij are complex numbers. Orthonormality of the new basis
vectors requires:
  
h10 |10 i = V11
∗ ∗
h1| + V12 h2| V11 |1i + V12 |2i (1.111)
† †
= V11 V11 h1|1i + V12 V21 h2|2i + . . . h1|2i + . . . h2|1i (1.112)
= 1. (1.113)

In the equations above we have used Vij† = Vji∗ , and we have not written explicitly
the coefficients in front of the terms h1|2i and h2|1i because both these terms vanish
due to the orthonormality of the vectors in B.
Going back to Eq. (1.112), and using h1|1i = h2|2i = 1, we obtain:

X
V1j Vj1 = 1. (1.114)
j=1,2

Similar relations can be obtained by considering the scalar products h10 |20 i, h20 |10 i,
and h20 |20 i. The four relations can be summarised as:

X
Vij Vjk = δik , for i, k = 1, 2 . (1.115)
j=1,2

This is nothing but the statement that Vij is a unitary 2 × 2 matrix. Eq. (1.109)
and Eq. (1.110) can be readily inverted:
† 0 † 0
|1i = V11 |1 i + V12 |2 i , (1.116)
† 0 † 0
|2i = V21 |1 i + V22 |2 i , (1.117)
and therefore
   
† † † †
|vi = v1 V11 + v2 V21 |10 i + v1 V12 + v2 V22 |20 i (1.118)
= v10 |10 i + v20 |20 i , (1.119)
which yields the coordinates of the same vector |vi in the new basis B 0 .
This is a simple 2 × 2 realization of the more general discussion that follows
Eq. (1.71). Check that you can match the two arguments. (Note that with the

notation used here we have Vkl = Ukl .)

Example 1.9 Two-state systems also provide a simple framework to construct


explicitly the tensor product of two vector spaces. Let us consider two spaces H1
and H2 , whose elements
 are the states of two independent two-state systems. Let
us denote by B1 = |1(1) i, |2(1) i and B2 = |1(2) i, |2(2) i their respective bases.
The set
n o
B = |1(1) 1(2) i, |1(1) 2(2) i, |2(1) 1(2) i, |2(1) 2(2) i (1.120)

is a basis of the four-dimensional vector space H = H1 ⊗ H2 . A generic vector in


H can be written as:
|ψi = c11 |1(1) 1(2) i + c12 |1(1) 2(2) i + c21 |2(1) 1(2) i + c22 |2(1) 2(2) i , (1.121)
xxii Quantum States

where c11 , c12 , c21 , c22 are complex coefficients. Setting

|V1 i = |1(1) 1(2) i , |V2 i = |1(1) 2(2) i , (1.122)


(1) (2) (1) (2)
|V3 i = |2 1 i , |V4 i = |2 2 i, (1.123)
C1 = c11 , C2 = c12 , C3 = c21 , C4 = c22 , (1.124)

allows us to rewrite the relation above as


4
X
|ψi = Ck |Vk i , (1.125)
k=1

which shows clearly that H is a four-dimensional vector space.


Let us now take one state |φ(1) i ∈ H1 and one state |φ(2) i ∈ H2 , these states
can be written as linear combinations of the basis vectors in their respective vector
spaces:
2 2
(1) (1) (1) (1)
|φ(1) i = c1 |1(1) i + c2 |2(1) i , c1 + c2 =1 (1.126)
2 2
(2) (2) (2) (2)
|φ(2) i = c1 |1(2) i + c2 |2(2) i , c1 + c2 = 1. (1.127)

We can build the tensor product of these two states, which is an element of H,

|φi = |φ(1) φ(2) i (1.128)


(1) (2) (1) (2) (1) (2) (1) (2)
= c1 c1 |1(1) 1(2) i + c1 c2 |1(1) 2(2) i + c2 c1 |2(1) 1(2) i + c2 c2 |2(1) 2(2) i .
(1.129)
The expansion of |φi is of the general form of Eq. (1.121). However, it is clear that
not all vectors in H can be written as a product like in Eq. (1.128).

Despite their apparent simplicity, two-state systems describe numerous physical


situations. They enter in the description of systems where only some degrees of
freedom are considered, e.g. when discussing the spin of the electron, or the polar-
ization of the photon, independently of other properties of these particles. Another
case where two-state system are relevant is when describing the transitions between
two atomic levels in a setting where other transitions are suppressed. Last but not
least, two-state systems are the building blocks of quantum computing, thereby pro-
viding the foundations for one of the most exciting new developments of Quantum
Mechanics.

1.3 The wave function of a one-dimensional system

Let us now consider the description of a quantum system in a one-dimensional space


– here we refer to actual space, so e.g. a point particle moving on a line, not to
xxiii The wave function of a one-dimensional system

the dimension of the Hilbert space of states. In this case, the state of the system is
described by the wave function, i.e. a complex function ψ(x) of the coordinate x
of the system. All the information about the state of the system is encoded in the
wave function. We need to reconcile the concept of a wave function with our claim
that quantum states are vectors in a Hilbert space.

1.3.1 Discretized System

In order to understand better how the information about the state of the sys-
tem is encoded in the wave function, we shall start with a simpler version of the
one-dimensional system, namely with a point particle in discretized space. Such a
particle can only be in a finite number of positions along the real axis, as shown
in Fig. 1.3. In this particular example, the particle can be in one of six 2 points
along the real axis labelled 0, . . . , 5. The lattice spacing (i.e. the distance between
two points) is denoted .

x
t
0 1 2 3 4 5

Fig. 1.3 Discretized one-dimensional system with six sites. The particle can only be in one of the six
points denoted by 0, . . . , 5. The distance between successive points is .

Let us assume that the state of the system is described by a wave function ψ;
then, at each point, xi , the wave function of the system takes some complex value
ψ(xi ). If we define

ψi = ψ(xi ), i = 0, . . . , 5 , (1.130)
then the state of the system is encoded in the values ψi of the wave function at the
discrete points xi , i.e. in a complex vector:
|ψi = (ψ0 , . . . , ψ5 ) . (1.131)

The reason for the factor of  in the definition of ψi will become clear below. The
values of ψ(x) at different spatial points are the coordinates of the state vector.
Two remarks are in order:
1. The vector |ψi in Eq. (1.131) is defined by giving its coordinates, and therefore
this definition assumes that a basis has been specified.
2. The x dependence is not explicitly written in the ket. Indeed the value of x
serves as a label for each coordinate of the vector, and in the discretised case it
is traded for the index i labelling the discrete set of points on the line. When
we refer to the vector we do not write the x dependence in the ket, just like
we do not add a coordinate index when we denote a vector by underlining its
corresponding symbol. We write v and not v i ; similarly we write |ψi and not
|ψ(x)i.
2 There is no particular significance in the fact that we have chosen here 6 points. The same
example could have been worked out with two points, or any finite number of points.
xxiv Quantum States

In the discretized system we see explicitly that the state of the system is rep-
resented by a vector. Because we considered a discretized space with six sites, we
obtain a six-dimensional complex vector. In the general case with N sites, we would
obtain an N -dimensional complex vector. This is an important concept to remem-
ber, as we will generalise it to the case of infinite-dimensional vector spaces.

Basis Vectors and Physical Interpretation


It is useful to introduce explicitly a basis such that the state vector |ψi can be
expanded as discussed above. The orthonormal base vectors are denoted

|ξ0 i = (1, 0, 0, 0, 0, 0) (1.132)


|ξ1 i = (0, 1, 0, 0, 0, 0) (1.133)
..
. (1.134)
|ξ5 i = (0, 0, 0, 0, 0, 1) . (1.135)

The vector |ψi can be written as


5
X
|ψi = ψk |ξk i , (1.136)
k=0

where we see that the complex numbers ψk are precisely the coordinates of |ψi in
the basis {|ξ0 i, . . . , |ξ5 i}. They are computed as usual by taking the scalar product
of the vector |ψi with the basis vectors

ψk = hξk |ψi . (1.137)

The physical interpretation of the basis vectors above is the following. Let us
assume that the modulus squared of the value of the wave function at a given
position x is the probability density of finding the system at x. We will discuss this
assumption in detail later when we discuss observables. In our discretized example,
2 √
|ψk | yields the probability of finding the system at xk . Note that the factor 
introduced in Eq. (1.130) is precisely the factor that is necessary to convert the
2
continuous probability density |ψ(x)| into a finite probability. The state described
by the basis vector |ξk i is a state in which the system has probability one of being
at position xk . For a generic state |ψi, the value of the wave function at xk , ψk , is
simply the projection of the ket describing the quantum state of the system onto a
basis vector that describes a state being localised at xk with probability one.
The scalar product of two generic vectors defined on the discretized line is:
X
hφ|ψi = φ∗k ψk ; (1.138)
k

once again notice that the coordinates of |φi are complex-conjugated when tak-
ing the scalar product. Following our convention, we write the norm as the scalar
xxv The wave function of a one-dimensional system

product of |ψi with itself:


5
X 2
||ψ||2 = hψ|ψi = |ψk | , (1.139)
k=0

and the normalization condition becomes


5
X 2
|ψk | = 1 . (1.140)
k=0

1.3.2 Continuum system

The wave function for a continuous system can be seen as the limit of the discretized
case where the number of points goes to infinity, while the distance  becomes
infinitesimally small. In this limit, instead of a finite-dimensional vector, we obtain
an infinite number of coordinates, encoded in a continuous function ψ(x).
We shall still refer to the wave function as the state vector, bearing in mind that
in this case the vector space is infinite-dimensional. Remember that the values ψ(x)
are interpreted as the coordinates of the vector |ψi: there is one coordinate ψ(x)
for each point x on the real axis, and there is an infinity of points along the real
axis. So the vector space is infinite-dimensional. We continue to denote the state
vector using a ket |ψi. It should be clear from the discussion above that the spatial
coordinate x labels the components of the state vector, and that the basis vectors
define a set of continuum states labeled by the variable x.
Interestingly, the concept of linear combination can also be extended to the case
of kets that are labeled by a continuous index x. First let us define the kets
1
|xk i = √ |ξk i . (1.141)

A linear combination of the |ξk i vectors can be written as:
X X√ √
ψk |ξk i = ψ(xk ) |xk i (1.142)
k k
Z
→0
= dx ψ(x) |xi ∈ H , (1.143)

where in the second line we have taken the limit  → 0, and we should remember
that ψ(x) is a complex function of x. The ket |xi is defined by taking the limit in
Eq. (1.141).
The completeness relation can be written as;
Z
dx |xihx| = 1 ; (1.144)

note that here we have a sum over an infinite number of projectors over the con-
tinuum basis vectors, we will come back to this point in the next chapter.
Similarly the norm of the state vector can be written as the limit of the norm of
xxvi Quantum States

the finite-dimensional vector for  → 0. Starting from Eq. (1.139), and taking the
limit:
X Z
2 2
hψ|ψi = lim  |ψ(xk )| = dx |ψ(xk )| , (1.145)
→0
k

where the sum over the discrete sites is replaced by the integral over the continuous
values of x. A state described by the wave function ψ is called normalizable, if
the integral in Eq. (1.145) is finite. More generally the scalar product of two wave
functions can be defined as the limit of the discrete case:
Z
hφ|ψi = dx φ(x)∗ ψ(x) . (1.146)

It is instructive to summarise these relations and compare them with the ones
we defined for discrete systems:

discrete system continuous system


P R
|ψi = ψ |ξ i R dx ψ(x)|xi
Pk k k
1= |ξ ihξ | dx |xihx|
Pk k 2 k 2
||ψ||2 =
R
|ψ | R dx |ψ(x)|
Pk ∗k
hφ|ψi = k φk ψk dx φ(x)∗ ψ(x)

Once we understand that the wave function encodes the set of coordinates of a
vector in an infinite-dimensional space, the connection with the previous discussion
about Hilbert spaces and the related tools becomes fairly natural.

Mathematical aside: square-integrable functions on the real axis.


2
The function ψ(x) = e−x /2 is an example of a normalizable function. Its norm is

Z
2
kψk2 = dx e−x = π , (1.147)

and therefore the normalized wave function is:


1
ψ(x) = exp[−x2 /2] . (1.148)
π 1/4
2
On the other hand, the function ex /2 is non-normalizable, and therefore does not
represent a physical state. The set of normalizable wave functions is the set of
square-integrable functions on the real line, denoted in the mathematical literature
as L2 (R).
2
In general, if dx |ψ(x)| = c, then the normalized wave function is √1c ψ(x).
R

It is interesting to consider what happens to the scalar product of two of the


basis states introduced in Eq. (1.132) ff. when we take the continuum limit, i.e. the
xxvii The wave function of a one-dimensional system

limit where  → 0:

1
hx|x0 i = lim hξ|ξ 0 i . (1.149)
→0 

Following the convention introduced earlier in this chapter the Greek letter ξ is used
to denote the states of the discretized system, while the Roman letter x stands for
the states in the continuum limit. We chose the states |ξi to be normalized to unity,
as discussed just above Eq. (1.132). Therefore hx|x0 i vanishes when x 6= x0 , while
it diverges for x = x0 . Using our definitions, we have

Z
hx|ψi = dx0 ψ(x0 )hx|x0 i (1.150)
X 1
= lim  ψ(ξk ) hξ|ξk i (1.151)
→0 
k
X
= ψ(ξk )δξ,ξk (1.152)
k

= ψ(x) . (1.153)

The scalar product hx|x0 i is the generalization of the Kronecker delta to the case of
continuum states, and is known as the Dirac delta. Some properties of the Dirac
delta are summarized below.

Mathematical aside: the Dirac delta


The delta function is the natural extension of the familiar Kronecker delta δij to
the case of continuous variables; it is defined as:

δ(x) = 0, if x 6= 0, δ(0) = ∞ , (1.154)

and

(
b
g(0), if a < 0 < b ,
Z
dx δ(x)g(x) = (1.155)
a 0, otherwise ,

for any continuous function g(x).


xxviii Quantum States

The following properties are often useful when dealing with delta functions.
Z b (
g(c), if a < c < b ,
dx δ(x − c)g(x) = (1.156)
a 0, otherwise ;
δ(−x) = δ(x) ; (1.157)
1
δ(ax) = δ(x) ; (1.158)
|a|
g(x)δ(x − y) = g(y)δ(x − y) ; (1.159)
xδ(x) = 0 ; (1.160)
(
d 1, if x ≥ 0 ,
θ(x) = δ(x) , where θ(x) = (1.161)
dx 0, if x < 0 ;
r
X 1
δ(g(x)) = δ(x − xi ) , where g(xi ) = 0 ; (1.162)
i=1
|g 0 (x i )|
Z ∞
1
dx g(x)δ(x) = g(0) . (1.163)
0 2
The delta function can be defined as the limit of ordinary functions. Here are some
examples; you are encouraged to sketch these functions in order to visualize what
happens when taking the limit:
1 2 2
δ(x) = lim √ e−x / , (1.164)
→0 π
sin(Lx)
= lim , (1.165)
L→∞ πx
sin2 (Lx)
= lim , (1.166)
L→∞ πLx2
1 
= lim . (1.167)
→0 π  + x2
2

The following integral representation for the delta function is very useful:
Z ∞
0
dx eix(k−k ) = 2πδ(k − k 0 ) . (1.168)
−∞

This completes the overview of the mathematical tools that will be used to char-
acterize and manipulate the states of a quantum system. They are the backbone of
our description of nature in the quantum realms. It is important to appreciate the
role of linear algebra and complex vector spaces in the description of physical states.
The physical states of every system described later in this book are described by
vectors in Hilbert spaces – make sure that you identify these vector spaces as you
read through the following chapters.
The next step in order to connect this abstract mathematical structure to the
physical world is the definition of a set of rules that allow us to deal with observables.
The next chapter is therefore devoted to the definition of observables in Quantum
xxix Problems

Mechanics and to the laws that ultimately will connect in a precise mathematical
way the results of experiments to the predictions of the theory.

Summary

• Quantum states are represented by vectors in Hilbert spaces. [1.1.1]


• The superposition principle emerges as a consequence of the mathematical
structure of vector spaces: a linear combination of state vectors with arbi-
trary complex coefficients yields a possible quantum state. [1.1.1]
• Functionals associate complex numbers to the vectors, and can be represented
by bras according to Dirac’s notation. [1.1.3]
• The scalar product in the Hilbert space allows us to define the norm of
the vectors. We will require that vectors representing physical states are
normalised to one. [1.1.4]
• Operators transform vectors into vectors. Acting with an operator on a quan-
tum state changes the state. [1.1.6]
• Vectors can be described by coordinates, after having chosen a particular
basis. The same vector is described by different coordinates in different
bases. [1.1.7]
• Given two vector spaces, it is possible to construct a new vector space, called
the tensor product of the two. Tensor products play a key role in the
description of composite systems. [1.1.8]
• The two-state system is a simple example of a complex vector space, which
describes the physical states of numerous real-life systems. The two-state
system provides simple explicit examples where we can practice with the
concepts introduced so far. [1.2]
• The wave function of a one-dimensional system can be viewed as a vector
in an infinite-dimensional vector space, thereby unifying the description of
quantum states as vectors. [1.3]

Problems

1.1 The energy levels for the infinite 1-dimensional square well are given by
~2 π 2 n 2
En = for n = 1, 2, 3, . . . ∞ . (1.169)
8ma2
Calculate the energies of the first (n = 1) and second (n = 2) levels for the case
xxx Quantum States

of an electron of mass 9.1 × 10−31 kg, confined to a box of atomic dimensions


(a = 10−10 m).
Calculate the wavelength of a photon emitted in a transition between these
levels.
1.2 Consider a three-dimensional vector space, and an orthonormal basis:
B = {|1i, |2i, |3i} .

Normalise the vectors

|vi = 2|2i ,
|φi = i|1i + 2|3i
|ψi = 2|1i + i|2i − 2 (cos α + i sin α) |3i ,

where α is a real number.


1.3 Given two orthonormal basis vectors |1i, |2i, consider a new set of basis vectors
1 i
|10 i = √ |1i + √ |2i
2 2
1 i
|20 i = − √ |1i + √ |2i .
2 2
Check that the new basis is orthonormal. Write down explicitly the 2 × 2
matrix that relates the coordinates of a vector in the basis {|1i, |2i} to the
coordinates of the same vector in the basis {|10 i, |20 i}, and check that it is a
unitary matrix.
The vector |vi has coordinates v1 = i, v2 = 3 in the basis {|1i, |2i}. Find
its coordinates in the basis {|10 i, |20 i}. Normalise the vector |vi.
1.4 The following inequalities recur frequently when working with vector spaces.
(a). Using the positivity of the norm of the vector |ψi − λ|φi, for all values of
the complex number λ, deduce the Schwarz inequality

|hψ|φi| ≤ ||ψ|| ||φ|| .

(b). Deduce the triangular inequalities, Eq. (1.37) and Eq. (1.38).
(c). Show that
||ψ + φ||2 + ||ψ − φ||2 = 2 ||ψ||2 + ||φ||2 .


1.5 The action of a linear operator U on the elements of a vector space H is


completely specified by the action of the operator on the vectors of one basis.
Let us consider a generic vector of a two-state system:

|ψi = ψ1 |1i + ψ2 |2i ,

then the action of U is given by

|ψi0 = Û |ψi = ψ1 Û |1i + ψ2 Û |2i .


xxxi Problems

Both Û |1i and Û |2i are elements of H, and therefore can be written as linear
combinations of the basis vectors:
|10 i = Û |1i = U11 |1i + U21 |2i
|20 i = Û |2i = U12 |1i + U22 |2i .

Find the coordinates ψ10 and ψ20 of the vector |ψ 0 i.


Show that, if the basis is orthonormal,
Uij = hi|Û |ji , i, j = 1, 2.

1.6 Consider two spaces H1 and H2 , whose elements are the states of two inde-
pendent
 (1) two-state subsystems
 denoted (1) and (2). Let us denote by B1 =
|1 i, |2(1) i and B2 = |1(2) i, |2(2) i their respective orthonormal bases.
The set
n o
B = |1(1) 1(2) i, |1(1) 2(2) i, |2(1) 1(2) i, |2(1) 1(2) i (1.170)

is a basis of the four-dimensional vector space H = H1 ⊗ H2 that contains the


states of the composite system made of the two subsystems. The subsystem
(1) is in the state
1 i
|ψ (1) i = √ |1(1) i + √ |2(1) i ,
2 2
and the subsystem (2) is in the state

(2) 1 (2) 2
|ψ i = √ |1 i + √ |2(2) i ,
3 3
find the coordinates of the state vector associated to the full system in the
basis B.
1.7 The action of a linear operator U on H1 and H2 is specified by the transfor-
mation properties of the basis vectors:
(1) (1)
|1(1)0 i = Û |1(1) i = U11 |1(1) i + U21 |2(1) i ,
(1) (1)
|2(1)0 i = Û |2(1) i = U12 |1(1) i + U22 |2(1) i ,
and
(2) (2)
|1(2)0 i = Û |1(2) i = U11 |1(2) i + U21 |2(2) i ,
(2) (2)
|2(2)0 i = Û |2(2) i = U12 |1(2) i + U22 |2(2) i .
Find the transformation properties of the vector
X
|ψi = ψij |i(1) i|j (2) i .
i,j=1,2

1.8 (a). Compute


Z +∞
dx sin(4x)δ(x − x0 ) .
−∞
xxxii Quantum States

(b). Compute
Z +∞
x
dx sin(x)δ( − x0 ) .
−∞ 4
(c). Compute
Z +∞
f (k) = dx eikx δ(x) .
−∞

1.9 The first and the third energy eigenstates of the one-dimensional harmonic
oscillator are respectively:
ψ0 (x) = C0 exp −α2 x2 /2 ,
 
(1.171)
ψ2 (x) = C2 4α2 x2 − 2 exp −α2 x2 /2 ,
  
(1.172)
2
where α = mω/~.
Calculate explicitly the normalization constants for these two energy eigen-
states, and verify that the eigenfunctions are orthogonal. Note that:
Z ∞ Z ∞
2 2 2 21 1√
exp{−α x }dx = (π/α ) and x2 exp{−α2 x2 }dx = π/α3
−∞ −∞ 2
(1.173)
1.10 Show that:
1 1 1   1
− = B̂ − Â . (1.174)
 B̂  B̂
Remember that  and B̂ are operators, and 1/ denotes the inverse of Â.
1.11 Given a sequence of vectors |vk i, we define a new set of vectors |uk i by the
following recursion relation:
1
1. |u1 i = |v1 i and |e1 i = p |u1 i.
hu1 |u1 i
k−1
X 1
2. |uk i = |vk i − |ej ihej |vk i and |ek i = p |uk i.
j=1
huk |uk i
Show by an explicit computation that |e1 i, |e2 i and |e3 i are orthonormal.
Show by induction that he1 |ek i = 0 for all k > 1.
This procedure is called Gram-Schmidt orthogonalization and is widely
used in constructing orthonormal bases.

You might also like