Handout Operator
Handout Operator
Operator methods in
quantum mechanics
Alongside the ket, we can define the “bra”, #ψ|. Together, the bra and ket
define the scalar product
! ∞
#φ|ψ" ≡ dx φ∗ (x)ψ(x) ,
−∞
from which follows the identity, #φ|ψ"∗ = #ψ|φ". In this formulation, the real
space representation of the wavefunction is recovered from the inner prod-
uct ψ(x) = #x|ψ" while the momentum space wavefunction is obtained from
ψ(p) = #p|ψ". As with a three-dimensional vector space where a · b ≤ |a| |b|,
the magnitude of the scalar product is limited by the magnitude of the vectors,
"
#ψ|φ" ≤ #ψ|ψ"#φ|φ" ,
3.1 Operators
An operator  is a “mathematical object” that maps one state vector, |ψ",
into another, |φ", i.e. Â|ψ" = |φ". If
Â|ψ" = a|ψ" ,
Moreover, for any linear operator Â, the Hermitian conjugate operator
(also known as the adjoint) is defined by the relation
! !
#φ|Âψ" = dx φ∗ (Âψ) = dx ψ(† φ)∗ = #† φ|ψ" . (3.2)
From the definition, #† φ|ψ" = #φ|Âψ", we can prove some useful rela-
tions: Taking the complex conjugate, #† φ|ψ"∗ = #ψ|† φ" = #Âψ|φ", and
then finding the Hermitian conjugate of † , we have
Therefore, if we take the Hermitian conjugate twice, we get back to the same
operator. Its easy to show that (λÂ)† = λ∗ † and ( + B̂)† = † + B̂ † just
from the properties of the dot product. We can also show that (ÂB̂)† = B̂ † †
from the identity, #φ|ÂB̂ψ" = #† φ|B̂ψ" = #B̂ † † φ|ψ". Note that operators
are associative but not (in general) commutative,
i.e. #Ĥψ|ψ" = #ψ|Ĥψ" = #Ĥ † ψ|ψ", and Ĥ † = Ĥ. Operators that are their
own Hermitian conjugate are called Hermitian (or self-adjoint).
' Exercise. Prove that the momentum operator p̂ = −i!∇ is Hermitian. Fur-
ther show that the parity operator, defined by P̂ ψ(x) = ψ(−x) is also Hermitian.
' Info. Projection operators and completeness: A ‘ket’ state vector fol-
lowed by a ‘bra’ state vector is an example of an operator. The operator which
projects a vector onto the jth eigenstate is given by |j"#j|. First the bra vector dots
into the state, giving the coefficient of |j" in the state, then its multiplied by the unit
vector |j", turning it back into a vector, with the right length to be a projection. An
operator maps one vector into another vector, so this is an operator. If we sum over
a complete set of states, like the eigenstates of a Hermitian operator, we obtain the
(useful) resolution of identity
&
|i"#i| = I .
i
%
Again, in coordinate form, we can write i φ∗i (x)φi (x" ) = δ(x − x" ). Indeed, we can
%
form a projection operator into a subspace, P̂ = subspace |i"#i|.
As in a three-dimensional
% vector %
space, the expansion of the vectors |φ" and
|ψ", as |φ" = i bi |i" and |ψ" = i ci |i",%
allows the dot product to be taken
by multiplying the components, #φ|ψ" = i b∗i ci .
' Example: The basis states can be formed from any complete set of orthogonal
states. In particular, they can' be formed from' the basis states of the position or
∞ ∞
the momentum operator, i.e. −∞ dx|x"#x| = −∞ dp|p"#p| = I. If we apply these
definitions, we can then recover the familiar Fourier representation,
! ∞ ! ∞
1
ψ(x) ≡ #x|ψ" = dp #x|p" #p|ψ" = √ dp eipx/! ψ(p) ,
−∞ ( )* +
√ 2π! −∞
eipx/! / 2π!
where #x|p" denotes the plane wave state |p" expressed in the real space basis.
Û = e−iĤt/! ,
denotes
% the time-evolution operator. By inserting the resolution of identity,
1
I = i |i"#i|, where the states |i" are eigenstates of the Hamiltonian with
eigenvalue Ei , we find that
& &
|ψ(t)" = e−iĤt/! |i"#i|ψ(0)" = |i"#i|ψ(0)"e−iEi t/! .
i i
1
This equation follows from integrating the time-dependent Schrödinger equation, Ĥ|ψ! =
i!∂t |ψ!.
p̂2
' Example: Consider the harmonic oscillator Hamiltonian Ĥ = + 12 mω 2 x2 .
2m
Later in this chapter, we will see that the eigenstates, |n", have equally-spaced eigen-
values, En = !ω(n + 1/2), for n = 0, 1, 2, · · ·. Let us then consider the time-evolution
of a general wavepacket, |ψ(0)", under
% the action of the Hamiltonian. From the equa-
tion above, we find that |ψ(t)" = n |n"#n|ψ(0)"e−iEn t/! . Since the eigenvalues are
equally spaced, let us consider what happens when t = tr ≡ 2πr/ω, with r integer.
In this case, since e2πinr = 1, we have
&
|ψ(tr )" = |n"#n|ψ(0)"e−iωtr /2 = (−1)r |ψ(0)" .
n
From this result, we can see that, up to an overall phase, the wave packet is perfectly
reconstructed at these times. This recurrence or “echo” is not generic, but is a
manifestation of the equal separation of eigenvalues in the harmonic oscillator.
' Exercise. Using the symmetry of the harmonic oscillator wavefunctions under
parity show that, at times tr = (2r + 1)π/ω, #x|ψ(tr )" = e−iωtr /2 #−x|ψ(0)". Explain
the origin of this recurrence.
Û † Û = I .
i #ψ|[Û , V̂ ]|ψ"
2λ(∆B)2 + i#ψ|[Û , V̂ ]|ψ" = 0, λ=− .
2 (∆B)2
i
∆A ∆B ≥ #[Â, B̂]" .
2
i !
∆p ∆x ≥ #[p̂, x]" = .
2 2
Similarly, if we use the conjugate coordinates of time and energy, [Ê, t] = i!,
we have
!
∆E ∆t ≥ .
2
d i, -
#ψ|Â|ψ" = #ψ|Ĥ Â|ψ" − #ψ|ÂĤ|ψ" +#ψ|∂t Â|ψ" .
dt (! )* +
i
#ψ|[Ĥ, Â]|ψ"
!
This is an important and general result for the time derivative of expectation
values which becomes simple if the operator itself does not explicitly depend
on time,
d i
#ψ|Â|ψ" = #ψ|[Ĥ, Â]|ψ" .
dt !
From this result, which is known as Ehrenfest’s theorem, we see that expec-
tation values of operators that commute with the Hamiltonian are constants Paul Ehrenfest 1880-1933
An Austrian
of the motion. physicist and
mathematician,
' Exercise. Applied to the non-relativistic Schrödinger operator for a single who obtained
2
p̂ %p̂& ˙ = −#∂x V ". Dutch citizenship
particle moving in a potential, Ĥ = 2m + V (x), show that #ẋ" = m , #p̂" in 1922. He
Show that these equations are consistent with the relations, made major
. / . / contributions
d ∂H d ∂H to the field
#x" = , #p̂" = − , of statistical
dt ∂p dt ∂x mechanics and its relations with
quantum mechanics, including the
the counterpart of Hamilton’s classical equations of motion. theory of phase transition and the
Ehrenfest theorem.
 → Û † ÂÛ .
Û † ÂÛ = R[Â] .
If Ô(p̂, r̂) ≡ Ĥ, the quantum Hamiltonian, such unitary transformations are
said to be symmetries of the quantum system.
one can show that, for a constant vector a, the unitary operator
# $
i
Û (a) = exp − a · p̂ ,
!
acting in the Hilbert space of a Schrödinger particle performs a spatial trans-
lation, Û † (a)f (r)Û (a) = f (r + a), where f (r) denotes a general algebraic
function of r.
where summation on the repeated indicies, im is assumed. Then, making use of the
Baker-Hausdorff identity (exercise)
1
e B̂e− = B̂ + [Â, B̂] + [Â, [Â, B̂]] + · · · , (3.3)
2!
it follows that
1
Û † (a)f (r)Û (a) = f (r) + ai1 (∇i1 f (r)) + ai ai (∇i1 ∇i2 f (r)) + · · · = f (r + a) ,
2! 1 2
where the last identity follows from the Taylor expansion.
As we have seen, time translations are generated by the time evolution op-
erator, Û (t) = exp[− !i Ĥt]. Therefore, every observable which commutes with
the Hamiltonian is a constant of the motion (invariant under time transla-
tions),
P̂ ψ(r) = ψ(−r) .
If we require that ψ(x, 2t) = ψ ∗ (x, 0), we must have Ĥ ∗ (x) = Ĥ(x). Therefore,
Ĥ is invariant under time-reversal if and only if Ĥ is real.
' Example: For example, if the Hamiltonian commutes with the angular mo-
mentum operators, L̂i , i = x, y, z, i.e. it is invariant under three-dimensional rota-
tions, an energy level with a given orbital quantum number , is at least (2, + 1)-fold
degenerate. Such a degeneracy can be seen as the result of non-trivial actions of
the operator L̂x and L̂y on an energy (and L̂z ) eigenstate |E, ,, m" (where m is the
magnetic quantum number asssociated with L̂z ).
Ĥ = !ω(n̂ + 1/2) .
1
|n" = √ (a† )n |0" ,
n!
The operators a and a† represent ladder operators and have the effect of
lowering or raising the energy of the state.
In fact, the operator representation achieves something quite remarkable
and, as we will see, unexpectedly profound. The quantum harmonic oscillator
describes the motion of a single particle in a one-dimensional potential well.
It’s eigenvalues turn out to be equally spaced – a ladder of eigenvalues, sepa-
rated by a constant energy !ω. If we are energetic, we can of course translate
our results into a coordinate representation ψn (x) = #x|n".7 However, the
operator representation affords a second interpretation, one that lends itself
to further generalization in quantum field theory. We can instead interpret
the quantum harmonic oscillator as a simple system involving many fictitious
particles, each of energy !ω. In this representation, known as the Fock space,
the vacuum state |0" is one involving no particles, |1" involves a single particle,
|2" has two and so on. These fictitious particles are created and annihilated
by the action of the raising and lowering operators, a† and a with canoni-
cal commutation relations, [a, a† ] = 1. Later in the course, we will find that
these commutation relations are the hallmark of bosonic quantum particles
and this representation, known as the second quantization underpins the
quantum field theory of the electromagnetic field.
' Info. There is evidently a huge difference between a stationary (Fock) state
of the harmonic oscillator and its classical counterpart. For the classical system, the
equations of motion are described by Hamilton’s equations of motion,
P
Ẋ = ∂P H = , Ṗ = −∂X H = −∂x U = −mω 2 X ,
m
where we have used capital letters to distinguish them from the arguments used to de-
scribe the quantum system. In the phase space, {X(t), P (t)}, these equations describe
a clockwise rotation along an elliptic trajectory specified by the initial conditions
{X(0), P (0)}. (Normalization of momentum by mω makes the trajectory circular.)
7
Expressed in real space, the harmonic oscillator wavefunctions are in fact described by
the Hermite polynomials,
r „r « » –
1 mω mωx2
ψn (x) = #x|n! = H n x exp − ,
2n n! ! 2!
2
dn −x2
where Hn (x) = (−1)n ex dxn
e .
On the other hand, the time dependence of the Fock space state, as of any sta-
tionary state, is exponential,
If we define  = α∗ a − αa† , then F̂α = e− and F̂α† = e . If we then take B̂ = I, we
have µ = 0, and F̂α† F̂α = I. This merely means that the shift operator is unitary -
not a big surprise, because if we shift the phase point by (+α) and then by (−α), we
certainly come back to the initial position.
If we take B̂ = a, using the commutation relations,
so that µ = α, and F̂α† aF̂α = a + α. Now let us consider the operator F̂α F̂α† aF̂α .
From the unitarity condition, this must equal aF̂α , while application of Eq. (3.4)
yields F̂α a + αF̂α , i.e.
Applying this equality to the ground state |0" and using the following identities,
a|0" = 0 and F̂α |0" = |α", we finally get a very simple and elegant result:
a|α" = α|α" .
8
After R. J. Glauber who studied these states in detail in the mid-1960s, though they
were known to E. Schrödinger as early as in 1928. Another popular name, coherent states,
does not make much sense, because all the quantum states we have studied so far (including
the Fock states) may be presented as coherent superpositions.
This means that the probability of finding the system in level n is given by the
Poisson distribution, Pn = |αn |2 = #n"n e−%n& /n! where #n" = |α|2 . More importantly,
δn = #n"1/2 / #n" when #n" 0 1 – the Poisson distribution approaches the Gaussian
distribution when #n" is large.
The time-evolution of Glauber states may be described most easily in the Schrödinger
representation when the time-dependence is transferred to the wavefunction. In this
case, α(t) ≡ √2x1
(X(t) + i Pmω
(t)
), where {X(t), P (t)} is the solution to the classical
0
equations of motion, α̇(t) = −iωα(t). From the solution, α(t) = α(0)e−iωt , one may
show that the average position and momentum evolve classically while their fluctua-
tions remain stationary,
3 41/2 3 41/2
x0 ! mωx0 !mω
∆x = √ = , ∆p = √ = .
2 2mω 2 2m
In the quantum theory of measurements these expressions are known as the “standard
quantum limit”. Notice that their product ∆x ∆p = !/2 corresponds to the lower
bound of the Heisenberg’s uncertainty relation.
' Exercise. Show that, in position space, the Glauber state takes the form
# $
mω Px
#x|α" = ψα (x) = C exp − (x − X)2 + i .
2! !