Sei sulla pagina 1di 6

Throughout this article, V is a vector space over the field of complex numbers

C, with the inner product h·, ·i. Absolutely no knowledge of physics is required,
but it is assumed that the reader is familiar with vector spaces and linear maps.
In mathematics, linear maps are commonly called linear transformations. In
physics, they are usually called linear operators. These three wordings are
equivalent.

1 Linear Operators
1.1 Hermitian operators
If a linear operator A over the space V has the following property:

hφ, Aψi = hAφ, ψi (1)


For all φ, ψ ∈ V , it is called Hermitian. As we will see, all transformations
corresponding to physical observables are Hermitian operators.

1.2 The Heisenberg Uncertainty Principle


Let A and B be two linear operators on V .
Proposition (The Uncertainty Principle) For some vector spaces, there exist
Hermitian operators A and B such that AB 6= BA.
Proof. This statement is so simple as to be almost trivial. A proof by
example is possible, and is left as an exercise for the reader.
Define the commutator of two linear operators as the following linear oper-
ator:

[A, B] = AB − BA (2)
Hermitian operators where [A, B] = i are said to be canonically conjugate.
Such pairs of operators are very important in quantum physics, and we will see
many examples of them later on.

1.3 Eigenvectors as bases


As we know, if A is a linear operator, and for some vector ψ and scalar a we
have:

Aψ = aψ (3)
Then ψ is called an eigenvector of A and a is the corresponding eigenvalue.
Now we are ready to prove a very useful fact about Hermitian operators:
Theorem. If ψ1 and ψ2 are two distinct eigenvectors of A and their corre-
sponding eigenvectors a1 and a2 are not equal, we have:

hψ1 , ψ2 i = 0 (4)
Proof. Since the two vectors are eigenvalues, we have:

hψ1 , Aψ2 i = hψ1 , a2 ψ2 i = a2 hψ1 , ψ2 i (5)

1
hψ2 , Aψ1 i = hψ2 , a1 ψ1 i = a1 hψ2 , ψ1 i (6)
Where we have used the property that the inner product is linear in the
second argument. Now we invoke the fact that A is Hermitian on (6):

hAψ2 , ψ1 i = a1 hψ2 , ψ1 i (7)


Take the conjugate:

hψ1 , Aψ2 i = a1 hψ2 , ψ1 i = a1 hψ1 , ψ2 i (8)


Where we used the property of conjugate symmetry of inner product. Now
subtract this from (5):

(a1 − a2 )hψ1 , ψ2 i = 0 (9)


If we consider the case where a1 = a2 , we see that since the dot product
on the right is nonzero, a1 = a2 . In other words, the eigenvalues of Hermitian
operators are invariant under conjugation. Now consider if a1 6= a2 . It is now
obvious that a1 6= a2 thus the dot product vanishes. Q.E.D.

2 Unitary transformations
Let U be a linear operator from V to the vector space W . If, for all vectors
in V , hU φ, U ψi = hφ, ψi then we say U is unitary. We call it anti-unitary if
hU φ, U ψi = hψ, φi.
Theorem (Wigner).

3 Quantum Mechanics
Quantum mechanics has a set of postulates. These are rules that have been
determined by experiment, and are given without proof. They relate our purely
mathematical world to the actual physical world.

3.1 Measurement
The most basic postulate of quantum mechanics is that the set of possible states
of a physical system is actually a subset of a vector space over a complex field,
and that the act of perfoming a measurement on the system is just equivalent
to transforming the state of the system with a Hermitian operator. In addition,
the eigenvalues of the operator are the possible values that the measurement
may give us.
Measurements include, for example, the position of a particle, the momen-
tum of a particle, etc.
One of the most important operators is the operator called the Hamiltonian.
The Hamiltonian is defined in such a way so that the eigenvalues correspond
to energy levels of a system. The precise definition, of course, depends upon
the problem in consideration, although it is usually defined as the sum of the
kinetic energy and potential energy operators of the system. The fundamental
postulate of quantum mechanics is that energy is quantized i.e. the set of energy
eigenvalues is countable, and that if a system is stationary (doesn’t change with

2
time) then the state of the system must be an eigenvector of the Hamiltonian.
This is put on a more formal footing below.

3.2 Schrödinger equation


Let H be the Hamiltonian of the system in question, and let Ψ be a map from
R to V (i.e. Ψ(t) is a member of V ) which is differentiable. The time-dependent
Schrödinger equation is:
d
−i Ψ = HΨ (10)
dt
According to quantum mechanics, if Ψ describes the time-evolution of a
system, then it must satisfy the above equation. Ψ is called the wavefunction.
Quantum mechanics is, at it’s core, the set of mathematical techniques that
is concerned with finding the solutions to the above equations, for different H
operators.

3.3 Multi-particle systems and quantum states


3.4 Fermions and Bosons
We have established that the states of a physical system live in a vector space.
The enquiring mind might ask: what of vector space isomorphisms? As it turns
out, isomorphisms — and, especially, automorphisms — play a central role in
quantum physics.
Let P be an automorphism on V . Consider what happens to the time-
independent Schrödinger equation if we transform it by P:
d
−iP Ψ = PHΨ (11)
dt
Due to the fact that isomorphisms have inverses, and P is independent of
time, we have:
d
−i PΨ = PHP−1 PΨ (12)
dt
So we see that if Ψ satisfies the wave equation with Hamiltonian H, then
PΨ satisfies it for the Hamiltonian PHP−1 . In particular, if P commutes with
the Hamiltonian (for example, multiplication with scalars), then PΨ satisfies
the equation for the same Hamiltonian.
As it turns out, in physical reality, just as vector space transformations
(operators) correspond to the act of observation, vector space automorphisms
that commute with the Hamiltonian correspond to the act of swapping particles.
That is, if we have a multi-particle system with wavefunction Ψ, then if we switch
it’s particles around, the wavefunction will be PΨ, for some P.
How do we know exactly which P to use? This is a question that can only
be determined by experiment.
To simplify matters, let’s just consider systems of two identical particles.
Experiment has determined that in this simple case, for some particle types we
simply have the trivial automorphism Pψ = ψ i.e. the wavefunction remains
the same upon switching particles. For others we have negation Pψ = −ψ i.e.

3
the wave function changes sign when the particles are interchanged. Particles
for which the former is true are called bosons. Particles for which the latter is
true are called fermions.
Experimental evidence has revealed that all particles are either fermions or
bosons1 2 .

3.5 Pauli exclusion principle


The peculiar property of fermions that their wavefunction changes sign upon
particle exchange immediately leads to the following observation. Let the wave-
function of a system of two identical fermions be ψ, and the wavefunction upon
0
their exchange be ψ . If the two fermions occupy the same quantum state then
0
if we switch them around their wavefunction must stay the same i.e. ψ = ψ
(Since have postulated that all the information about a system is captured by
it’s wavefunction). On the other hand, we know that upon the exchange of
0
fermions their sign is switched so ψ = −ψ . This leads to ψ = 0. Clearly
a wavefunction that is zero does not make physical sense. This paradoxical
situation forces us to conclude that :
Proposition (exclusion principle): No pair of identical fermions may occupy
the same quantum state.

3.6 Examples
So far we have been talking about states and Hamiltonians, slyly avoiding any
actual concrete examples of what they might be. We are now ready to in-
troduce some concrete examples of quantum systems and their corresponding
Hamiltonians.

4 Bra-ket notation
The bra-ket notation is a peculiar notation used by physicists for dealing with
linear operations on spaces that have certain properties. It may be used to
simplify many calculations in quantum physics.
In the following, this notation is given a relatively solid mathematical footing.
We have deviated slightly from the usual physical notation. Specifically, kets
are defined differently. In physics, kets are simply vectors. Here, they are
functionals. If we place certain constraints on V , these two definitions will be
seen to be functionally equivalent. The reason for proceeding like this will be
seen later on.
1 There has also been hypothesized a third kind of particle — so-called anyons — whose

automorphisms are given by Pψ = zψ, where z could be any complex number. These particles
are the basis of topological quantum computing. However, in 3 dimensions such particles have
not been shown to exist.
2 Note that we did not say ‘elementary particles’. In fact, all objects, including macroscopic

objects such as human beings, are either fermions or bosons. However, at larger scales fermions
and bosons become almost indistinguishible due to the fact that many energy states become
available and no two objects will ever really occupy the exact same one. Even molecules are
too large for bosonic or fermionic effects to reveal themselves, except at very low temperatures.
At such low temperatures, it is possible to group large numbers of bosons together into a single
quantum state. Such an object is called a Bose-Einstein condensate.

4
4.1 Bras and kets
Let ψ be a vector in V and define the ket |ψi as:

|ψi :V → F
|ψi :φ 7→ hφ, ψi

(the observant reader will have realized that |ψi is a linear functional).
Similarly, define the bra hφ| as:

hφ| :V → F
hφ| :ψ 7→ hφ, ψi

4.2 Riesz Representation Theorem


This theorem states that if V is a Hilbert space over the complex numbers, and
f is a functional over V , then we can find a unique φ ∈ V such that f = hφ|
and a unique ψ ∈ V such that f = |ψi.
Proof. TODO.
The theorem says that, in a sense, |ψi and ψ can be considered to be ‘inter-
changeable’.
With this theorem in hand, we can go on to some useful definitions.

4.3 Inner and Outer product


Let V be a Hilbert space over the complex numbers as described above, and
let ϕ and ϕ∗ be the functions that map functionals to their unique kets and
bras, respectively. Due to the Riesz theorem, these functions are well-defined.
Obviously, ϕ(|ψi) = ψ and ϕ∗ (hφ|) = φ.
Let f and g be two functionals over V . The inner product f.g is defined as
hϕ∗ (f ), ϕ(g)i, and the outer product f × g is defined as:

f × g : h 7→ hϕ∗ (g), ϕ(h)if (13)


Especially, |ψi × hφ| : |ρi 7→ hφ, ρi|ψi.
At the risk of being redundant, it is important to note that the inner product
is a (complex) number and the outer product is a linear operator on functionals.
When dealing with bras and kets, wherever there is no ambiguity, hφ|ψi will
be taken to mean the inner product of the functionals hφ| and |ψi, and |ψihφ|
will be taken to mean the outer product (i.e. |ψi × hφ|, not hφ| × |ψi). To clarify,
we use the relative positions of the brackets to signify what type of product we
are using.
Now, using this notation, we can define an ‘algebra’ on bras and kets. For
example, take hφ|ψihρ|. The key observation that we can make is that the two
interpretations (namely, (hφ|ψi)hρ| and hφ|(|ψihρ|)) are equal (Verify this).
Another interesting observation is that hφ|ψ = hφ|ψi. In other words, the
action of the functional hφ| on ψ is the same as the dot product of φ and
ψ. In fact, in physics, these two forms are taken to be equal and are used
interchangeably.

5
To complete our exposition of the bra-ket notation, we introduce linear op-
erators into the mix. A linear operator on V is, of course, not defined on
functionals of V . However, if the representation theorem applies to V , then we
can make such a definition. Let f be a functional on V and let A be a linear
operator on V . Then define the following operation:

A(f ) = Af = |Aϕ(f )i (14)


The full bra-ket notation machinery is now defined. Using what has been
discussed, the following follow immediately:

1. A|ψi = |Aψi
2. hφ|A|ψi = hφ|Aψi
3. hφ|I|ψi = hφ|ψi

4.4 Examples
(Finite dimensional, Hilbert spaces, L2 , etc.)
In quantum mechanics, the 3 vector spaces that we use most often are
the finite-dimensional vector space over the complex numbers, the infinitite-
dimensional Hilbert space over the complex numbers, and the vector space of
square-integrable complex functions L2 .
In the space of square-integrable functions on the vector space S defined
over the complex field, with the measure µ, define the inner product:
Z
hφ, ψi = φ(x)ψ(x) dµ(x). (15)
S
In particular, for complex square-integrable functions over the real line, we
have:
Z ∞
hφ, ψi = φ(x)ψ(x) dx. (16)
−∞

The representation theorem holds for both of these spaces.

Potrebbero piacerti anche