You are on page 1of 7

Postulates of Quantum Mechanics

Postulate 1. (The trivial postulate) The state of a quantum mechanical


system is completely specified by a Hilbert space vector (r, t), known as the
wavefunction. The wave function must be continuous and finite everywhere, and
must satisfy the following condition:

Where d is a volume element located at r.

Postulate 2. (The Probabilistic Law) This law has three parts:


I.
II.

III.

When an observable A(P, r, E) is measured, one of the eigenvalues of A(


, , ) (a Hermitian operator) is obtained.
The state vector can be expanded with the eigenvectors of A as a basis.
The probability of obtaining a particular eigenvalue as the result of
measurement is given by the squared of the modulus of the coefficient of
the corresponding eigenvector in the expansion of the total state vector.
The act of measurement causes the wavefunction to collapse to the
corresponding eigenvector (or to the projection of onto the degenerate
subspace, in case the eigenvalue is degenerate).

Table of physical observables and their corresponding quantum operators (single


particle)
Observable

Observable

Operator

Operator

Name

Symbol

Symbol

Operation

Position
Momentum

Multiply by

Total energy

Dynamic variable

A(r, p, E)

A( , ,

Hence, the average value of the observable corresponding to

is given by

From the law of conservation of energy, we know that the observable f = T + V E


= 0. This implies that the average value of f, <f>, equals zero. Hence, f=0. Hence,
(T + V E) = 0. This is called the Schrodingers equation. The reason why this
equation works only in non-relativistic limits is that the expression of T used here
is non-relativistic.

A discussion on the postulates:


One can mathematically show that for any continuous linear operator A: H H,
there exists a unique continuous linear operator A*: H H with the following
property:

The operator A* is called the adjoint of A. If A* = A, then the operator A is called a


Hermitian operator. In case A* = A-1, the operator A is called a Unitary operator.
Hermitian operators always have real eigenvalues, which is consistent with the fact
that measurements always return one of the eigenvalues of the operator
corresponding to the observable.
With regard to the third postulate, it is interesting to note that we can only know
the probability of finding the particle in a particular state, and this is not due to the
inadequacy of quantum mechanics. The reason we can only know about the

probability that a measurement will return a particular value is due to the fact that
the various dynamic variables which we employ in order to study microscopic
particles do not have well-defined values. In short, the particle doesnt have a welldefined state before the measurement occurs. Let us take an example where we are
measuring the position of the particle. The particle doesnt have an exact position
until the position is measured. The act of measurement forces the particle to
randomly assume some position, with the probability density given by the modulus
of the wave function. This event is called the wave function collapse, because
after the measurement, the wave function collapses to become peaked at the
measured value.
Thus, there are two different kinds of physical processes which occur in quantum
mechanics: the ordinary ones, in which the state of a system evolves undisturbed
according to some wave equation (the Schrdinger equation is a good
approximation for non-relativistic limits), and measurements, which cause the
wave function to suddenly and discontinuously collapse. The mechanism of
collapse is unknown, and the word measurement yet lacks a rigorous definition.
Moreover, since the particle doesnt have a well-defined position before
measurement, we can no longer give meaningful physical interpretations to terms
like velocity, momentum, trajectory of the particle, etc. In fact, these terms are not
well-defined for microscopic particles, because such particles simply dont have a
well-defined position, momentum, etc. before they are measured.
To make the matters worse, it turns out that there are certain pairs of conjugate
variables, like position and momentum, which cannot be simultaneously
measured. It is engrained within the very nature of the wave function that it cannot
collapse into a state which simultaneously has an exact position and momentum.
Thus, if we know with certainty about the exact position of a particle at some
instant, the particle simply wont have a well-defined momentum, such that if the
very next moment its momentum were measured, it could turn out to be anything.
The more precisely the position of some particle is determined, the less precisely
its momentum can be known, and vice versa. A formal inequality relating
the standard deviation of position x and the standard deviation of momentum p is
given as:

Particles in quantum mechanics have a quantum-mechanical property called spin.


Spin is an intrinsic form of angular momentum carried by elementary particles,
composite particles (hadrons), and atomic nuclei. As the name suggests, spin was
originally conceived as the rotation of a particle around some axis. This picture is
correct so far as spin obeys the same mathematical laws as quantized angular
momenta do. On the other hand, the spin of an elementary particle is a truly
intrinsic physical property, akin to the particle's electric charge and rest mass.
Theoretical and experimental studies have shown that the spin possessed by such
particles cannot be explained by postulating that they are made up of even smaller
particles rotating about a common center of mass; as far as can be determined,
these elementary particles have no inner structure.
When the idea of electron spin was first introduced in 1925, even Wolfgang Pauli
had trouble accepting Ralph Kronig's model. The problem was not that a rotating
charged particle would have given rise to a magnetic field but that the electron was
so small that the equatorial speed of the electron would have to be greater than the
speed of light for the magnetic moment to be of the observed strength.
In 1930, Paul Dirac developed a new version of the Wave Equation which
was relativistically invariant (unlike Schrdinger's one), and predicted the
magnetic moment correctly, and at the same time treated the electron as a point
particle. In the Dirac equation all four quantum numbers including the additional
quantum number, s arose naturally during its solution.
Identical particles, also called indistinguishable or indiscernible particles,
are particles that cannot be distinguished from one another, even in principle.
Species of identical particles include elementary particles such as electrons, and,
with some clauses, composite particles such as atoms and molecules.
There are two ways in which one might distinguish between particles. The first
method relies on differences in the particles' intrinsic physical properties, such
as mass, electric charge, and spin. If differences exist, we can distinguish between

the particles by measuring the relevant properties. However, it is an empirical fact


that microscopic particles of the same species have completely equivalent physical
properties. For instance, every electron in the universe has exactly the same
electric charge; this is why we can speak of such a thing as "the charge of the
electron".
Even if the particles have equivalent physical properties, there remains a second
method for distinguishing between particles, which is to track the trajectory of each
particle. As long as we can measure the position of each particle with infinite
precision (even when the particles collide), there would be no ambiguity about
which particle is which.
The problem with this approach is that it contradicts the principles of quantum
mechanics. According to quantum theory, the particles do not possess definite
positions during the periods between measurements. Instead, they are governed
by wavefunctions that give the probability of finding a particle at each position. As
time passes, the wavefunctions tend to spread out and overlap. Once this happens,
it becomes impossible to determine, in a subsequent measurement, which of the
particle positions correspond to those measured earlier. The particles are then said
to be identical, or indistinguishable.
Two indistinguishable particles only have one state, not two. This means that if we
exchange the positions of the particles, we do not get a new state, but rather the
same physical state. In fact, one cannot tell which particle is in which position.
A physical state is described by a wavefunction, or more generally by a vector,
which is also called a "state"; if interactions with other particles are ignored, then
two different wavefunctions are physically equivalent if their absolute value is
equal. So, while the physical state does not change under the exchange of the
particles' positions, the wavefunction may get a minus sign.
Bosons (particles with whole number spin) are particles whose wavefunction is
symmetric under such an exchange, so if we swap the particles the wavefunction
does not change. Fermions (particles with half-integer spin) are particles whose
wavefunction is antisymmetric, so under such a swap the wavefunction gets a
minus sign, meaning that the amplitude for two identical fermions to occupy the

same state must be zero. This is the Pauli Exclusion Principle: two identical
fermions cannot occupy the same state. This rule does not hold for bosons.
Another strange aspect of quantum mechanics is that although the position of a
particle is not certain, the time recorded by its internal clock does not suffer from
the same uncertainty. In short, particles dont have real positions, but they do have
a real age. If this were not the case, then we would have to discard our usual
approach of treating time as a parameter, and without a well-defined parameter we
would not be able to describe the evolution of physical systems. The whole of
physics would be thrown into disarray, and we would have no clue so as to which
direction we should advance. Thankfully, that didnt happen.
What quantum mechanics currently lacks is a proper system to determine the
operator for any observable, given the three basic operators. It also lacks a
theoretical justification of the exact relationship between operators and
observables. What is the theoretical connection between a dynamical variable, and
its quantum operator?
One has an intuitive feeling that the classical definition of the dynamical variables
and the mathematical nature of the operators have some connection. If we
understood this connection, we would also understand why the eigenvalues of the
operators correspond to the possible outcomes of measurement. This could also
have got something to do with the very nature of the wavefunction.
What we need is the correct mathematics that would enable us to easily construct
operators, given any observable. I can see half of the problems melting away if we
could do that. For example, how would you construct the operator for XP (position
times momentum)? Is it XP or PX (since the two operators do not commute)? This
ambiguity cannot be theoretically resolved, and only an experiment can provide an
answer. For another instance, using the usual method of operator construction does
not enable us to construct an operator for the rate of change of momentum with
respect to time (because the rate of change of the momentum operator with respect
to time is simply meaningless; operators dont change with time).
In short, we need the exact science of operator construction. Since quantum
particles dont have a real position, the phrase, measuring the position is actually
misleading. A more accurate description of the measurement process would be like

this: when we interact with the quantum particle using the same procedure as used
in measuring the position of a classical particle, the value we get is one of the
eigenvalues of the position operator. The measurement procedure has a direct
effect on the wavefunction, and the procedure itself is related to the classical
definition of the observable. Hence, the nature of the observable and the effect it
has on the wavefunction must have some connection. It is my intuitive guess that
all the answers lie in the nature of the wavefunction, including the way it interacts
with its surroundings.
We know that when an observable is a linear symmetric combination of the three
basic observables, we can find its operator simply by replacing the three basic
variables by their respective observables. However, if the combination is linear but
not symmetric, we get an ambiguous case, and if the combination is not linear, we
get meaningless results when we apply the above method for operator construction.
For instance, what would be the operator for sin P? Supposedly, the sine of the
momentum operator is supposed to be another operator the one corresponding to
sin P. Basically what this means is that if A is an operator corresponding to A, then
f(A) is an operator whose eigenvalues are the f of the eigenvalues of A. This
definition would then be consistent with the condition that a measurement of A
returns one of the eigenvalues of A. So if we measure f(A), the result should be the
f of the eigenvalues of eigenvalues of A. However, these values are precisely
supposed to be the eigenvalues of f(A).

You might also like