You are on page 1of 27

Notes on PSO201A (Quantum Physics; 2019-20-II)

Anjan K. Gupta, Department of Physics, IIT Kanpur

1 Linear Vector Spaces (LVS):


Let’s start with a concrete example of linear vector space (LVS) that most of us are
already familiar with, i.e. the three dimensional LVS of vectors. We know how a vector
is an abstract entity and we choose a set of axes, or bases, in order to work with vectors
mathematically. The exact mathematical form of a given vector depends on the bases (or
axes) that we choose in 3D space. These same concepts will be generalized for quantum
mechanics as LVS play a very important role in QM. To define a LVS formally: it is a set

→ − → − → − →
of elements, say {V1 , V2 , V3 , V4 . . . }, satisfying the following

→ − → −
→ −

1. If V1 & V2 are two vectors, i.e. elements of this set, than λ1 V1 +λ2 V2 is also a vector.
Here λ1 , λ2 are real numbers and so this LVS is called real LVS. We similarly have
complex LVS also.

→ − →
2. There is a scalar product defined, i.e. V1 · V2 , which is a scalar, i.e. a real number
in case of real LVS. We know what this scalar product means for a LVS of vectors,
i.e. it is the scalar or dot product of two vectors. Also this scalar product satisfies
a few properties, namely,

→ − →
(a) The scalar product is linear with respect to both V1 & V2 , i.e.
−→ −
→ − → −→ − → −
→ − →
(λ1a V1a + λ1b V1b ) · V2 = λ1a V1a · V2 + λ1b V1b · V2 . (1)


Similarly it is linear wrt V2 .

→ − → −
→ −
→ −
→ −

(b) If V1 · V2 = 0 than either V1 or V2 is zero, i.e. null vector, or V1 & V2 are
orthogonal, i.e. mutually perpendicular in case of vectors in 3D space.

→ − → −

(c) V · V = 0 if and only if V is a null vector. The norm of the vector, i.e.

→ −
→ −
→ −
→ − →
|| V || (or simply | V |), gets defined as || V ||2 = V · V . Norm represents the
magnitude of the vector and it is a positive number.
(d) Schwarz inequality: The scalar product satisfies

→ − → −
→ −

(V1 · V2 )2 ≤ ||V1 ||2 ||V2 ||2 . (2)

→ − → −
→ − →
Here the equality holds if V1 ∝ V2 , i.e. if V1 & V2 are parallel or co-linear.

1
With the scalar product defined one can also define the angle between two vectors and this
can be generalized to other LVS as well. These vector spaces are also called Hilbert space,
which is a common terminology for any quantum system as the Hilbert space encompasses
all possible states wrt to a particular degree of freedom such as the spin degree of freedom
or spatial (3D) degree of freedom.

1.1 Bases Vectors:


In the LVS of vectors in 3D a vector (an abstract entity) is mathematically written on
paper after choosing a set of axes and the bases vectors along these axes, i.e. ub1 , ub2 ,
ub3 . These satisfy the orthonormality, i.e. orthogonality & normalized to one, condition,
written as,

ubi · ubj = δij. (3)




with δij as the Kronecker delta, i.e. δij = 1 if i = j and zero otherwise. Now any vector, V

→ P
can be written as V = V1 ub1 + V1 ub2 + V1 ub3 = i V1 ubi . Representability of any vector (or
element of the LVS) in terms of ubj is very important. It is also called completeness of the
bases or the closure condition. Here Vi is a number and it represents the ith component


of V . These {Vi }, i.e. their values, depend on which bases we choose. In 3D we are free
to choose any set of three orthogonal axes and corresponding unit vectors as bases. The
number of independent basis vectors required for completeness is called the dimensionality
of the LVS, which is 3 for LVS of vectors in 3D.


After choosing this representation (or bases) a particular component of V can be
written as linear superposition of components along the basis vectors, which are the scalar


products with corresponding basis vector, i.e. Vi = V · ubi . In terms of these components,
We can write the scalar product between two vectors as

→ − → X X X
U ·V = Ui ubi · Vj ubj = Ui Vj δij = Ui Vi . (4)
ij ij i


→ −
→ −
→ − →
Furher, if we look at U and V as two column matrices than the scalar product U · V
→ −
− →
is equivalent to the matrix product ( U )T V where the first one is the transpose of the


column matrix U giving a row matrix. This’ll also help us appreciate the Dirac notation
discussed later. This can be looked at more explicitly as,
 

→ − → X  V1 −
→ − →
U ·V = Ui Vi = U1 U2 U3 V2  = ( U )T V
i V3

While we all are familiar with above stuff for LVS of vectors in 3D space, it’s useful to
keep this in mind as a specific example of LVS. This gives some intuition and and it’s good
for remembering certain relations for other LVS that are more complex and indispensable
in QM.

2
1.2 Linear Operators:
Again we start with operators in 3D LVS of vectors. We also have come across some
operators such as rotation operators, moment of inertia tensors that map a given vector
on to another one. Afterall an operator defines a one to one mapping from one element of
LVS on to another. So the rotation operator gives the vector after rotating a given vector
by certain angle about certain axis. Naturally this is a one-to-one mapping and it also
satisfies a linearity. Operators are also abstract entities that can be written concretely
after choosing a bases and its explicit mathematical form, just like vectors, will depend
on the chosen bases. One can of course find how an operator (or vector) will change once
we go from one basis to another. We’ll see this later.
A few properties satisfied by these linear operators are:

→ − → −
→ − → −
→ −

1. Linearity: If A is an operator so that AV1 = V1 ′ & AV2 = V2 ′ then A(λ1 V1 +λ2 V2 ) =

→ −

λ1 V1 ′ + λ2 V2 ′ .

2. Operators’ product satisfies the associative law, i.e. (AB)C = A(BC) but not the
commutative law (in general), i.e. AB ̸= BA. There may be some special operators
that commute with each other and this is of special significance in QM.

To write an operator A in concrete mathematical form using certain bases {ubi } we analyze

→ → P
− −
→ P
its operation on a given vector V in the same bases. Suppose V = Vi ubi , V ′ = Vi′ ubi

→ − → −

and A V = V ′ . Taking scalar product of the last relation with ubj , we get ubj · (A V ) =

→ P P
ubj · V ′ = Vj′ . This gives Vj′ = ubj · (A i Vi ubi ) or Vj′ = i (ubj · Aubi )Vi . Thus we see that we
can find the result of A operating on any given vector if we know the entities {ubj · Aubi }
corresponding to a given operator P A. These entities are called matrix elements of A, i.e.
Aji = ubj · Aubi and so we get Vj′ = i Aji Vi .
It is more convenient to work with matrix form of operators and vectors. With the

→ −

matrix elements defined above, we can write V ′ = A V as
 ′   
V1 A11 A12 A13 V1
V2′  = A21 A22 A23  V2  (5)
V3′ A31 A32 A33 V3

The above defines the matrix forms of vectors, i.e. elements of LVS, and operators after
choosing a bases, i.e. a complete set of basis vectors. One can also define general matrix

→ − →
element of anPoperator as V1 · AV2 which can be written in matrix form and eventually it
simplifies to ij V1j Aji V2i . We should realize that the “linearity” of operators and vectors
plays an important role in all these simplifications.

1.3 Change of bases:


Change of bases is quite important in QM since finding the eigen-values and eigen-states
of an operator, such as Hamiltonian, is basically finding a bases in which the operator is
diagonal. We emphasize that the change of basis does not actually change the state or the
operator physically but its appearance changes. We consider two different orthonormal
bases {ubi } and {vbi }, which in case of LVS of vectors in 3D can be associated with two

3
different set of orthogonal axes that are rotated wrt each other. Given a vector or operator
in one bases our objective here is to find them in the other bases.


Suppose in {ubi } bases a vector V has components {Vi } while in {vbi } bases the com-
ponents are {Vi′ }. We have to find how the two components are related to each other.
Clearly the two represent the same vector P but inPtwo different bases. Writing the two

forms of the same vector P as equal, i.e. V b
u
i i i = j Vj vbj , and taking its scalar product
with vbk we get = Vk′ = i Vi vbk · ubi . Defining

Ski = vbk · ubi (6)


P
we get the linear relation Vk′ = i Ski Vi . Thus we see that

→′ −

V = SV (7)

gives the vector in the new basis. We can see that for real LVS the transformation matrix
S is a real matrix with elements Ski = vbk · ubi as the scalar products between the elements
of the two bases. Given the orthonormality of the two bases, we can also verify that SS T
is a unit matrix,
P T i.e.′ S is unitary. Thus one can also write the inverse transformation,
T
i.e. VPk = i Ski Vi . Here S is the transpose of S. Using this S we can also write
vbk = j ubi (ubi · vbk ), i.e.
X
vbk = Ski ubi . (8)
j

Similarly for operator A, in {ubi } bases its matrix elements are {Aij } while in {vbi }

bases these are {Aij }. We have to find how the P two are related
P to each P other. For this

we use the above Eq.8 to write, Akl = vbk · Ab vl = i Ski ubi · A j Slj ubj = ij Ski Aij Slj =
P
ij Ski Aij Sjl . Thus in the {v
bi } bases the operator A will be given by
T

A′ = SAS T . (9)

We see that the change of bases is captured, for both the vectors and operators, by
the unitary matrix S and thus this is also called a unitary (or similarity for real LVS)
transformation.

→ −

We can also see that a scalar expression, such as V1 · AV2 works out same no mat-
ter which bases is chosen. For this we need to look at the scalar product in the ma-

→ −
→ −
→ −

trix form. V1 · AV2 in the second bases will be V1 ′ · A′ V2 ′ , which using matrix notation

→ −

will give (V1 ′ )T A′ V2 ′ . Using the change of bases transformation matrix S, this gives

→ −
→ −
→ −
→ − → − → − → −

(S V1 )T (SAS T )(S V2 ) = V1 T S T SAS T S V2 = V1 T AV2 = V1 · AV2 . This uses unitary prop-
erty of S. This illustrates that such scalar expressions are independent of bases. It is
useful to work out some such examples to gain some experience.

2 Complex LVS of wave-functions:


This is a LVS of the well behaved, i.e. normalizable and analytical, complex wave-
functions. Unlike the LVS of vectors in 3D, where any vector can be written in terms

4
of three independent numbers, this LVS is infinite dimensional. Thus the set of required
basis wave-functions (or also called basis vectors) is infinite. We can re-state the definition
of this LVS as there are differences in detail:
1. If ψ1 (x) and ψ2 (x) are two valid wave-functions, i.e. elements of this LVS, than
c1 ψ1 (x)+c2 ψ2 (x) is also a valid wave-function. Here c1 and c2 are complex numbers.

2. The scalar product is defined as


Z ∞
⟨ψ1 |ψ2 ⟩ = ψ1∗ (x)ψ2 (x)dx. (10)
−∞

The notation used on left for scalar product is called Dirac notation and we’ll discuss
this in more detail later.

(a) Clearly, this scalar product is linear wrt ψ2 and it is anti-linear wrt ψ1 , i.e.
⟨(c1a ψ1a + c1b ψ1b )|ψ2 ⟩ = c∗1a ⟨ψ1a |ψ2 ⟩ + c∗1b ⟨ψ1b |ψ2 ⟩ and ⟨ψ1 |(c2a ψ2a + c2b ψ2b )⟩ =
c2a ⟨ψ1 |ψ2a ⟩ + c2b ⟨ψ1 |ψ2b ⟩
(b) If ⟨ψ1 |ψ2 ⟩ = 0 then either ψ1 (x) of ψ2 (x) is zero or ψ1 (x) and ψ2 (x) are orthog-
onal.
(c) The norm of a wave function ψ(x) gets defined using the scalar product as
Z ∞
||ψ|| = ⟨ψ|ψ⟩ =
2
|ψ(x)|2 dx. (11)
−∞

Thus the norm is real and positive with well defined square-root, which gives
the overall probability in QM.
(d) The scalar product satisfies the Schwarz inequality, i.e. |⟨ψ1 |ψ2 ⟩|2 ≤ ||ψ1 ||2 ||ψ2 ||2

One can easily generalize the notion of linear operators to this LVS as well. An operator
is still a linear mapping from one wave-function to another, i.e. ϕ(x) = Aψ(x). The
linearity implies A[c1 ψ1 (x) + c2 ψ2 (x)] = c1 [Aψ1 (x)] + c2 [Aψ2 (x)]. One can re-look at
this with already encountered operators, such as position operator (x), p, H, etc. A few
others are, parity operator, defined by Πψ(x) = Ψ(−x) or translation by a operator, i.e.
Ta ψ(x) = ψ(x + a). The product of operators are also easily generalized. The products
again follow the associative law but not the commutative law in general. The later is
important for QM and one defines the commutator between two operators A & B as
[A, B] = AB − BA; for eg. [x, px ] = iℏ.

2.1 Dirac notation:


The quantum mechanical state, specified by a wave-function, of a system is an abstract
entity and once we put it down on paper as wave-function we have already chosen a bases.
That bases is x-basis, i.e. {|x⟩}. We have seen that we can represent the same quantum
state in p-basis, i.e. {|x⟩} by using the Fourier transform (FT) as
Z ∞
1
gψ (p) = √ ψ(x) exp(−ipx/ℏ)dx (12)
2πℏ −∞

5
or we can transform it back using inverse Fourier Transform (IFT) as
Z ∞
1
ψ(x) = √ gψ (p) exp(ipx/ℏ)dp. (13)
2πℏ −∞
Thus we can choose a representation (or bases) to write the abstract state in mathematical
form. Dirac introduced a notation to handle these abstract states without choosing a
specific bases. This notation consists of three expressions: 1) |ψ⟩, pronounced as ket-ψ,
2) ⟨ϕ|, i.e. bra-ϕ, and 3) ⟨ϕ|ψ⟩, i.e. bra-ϕ-ket-ψ. The last quantity is the scalar product, in
abstract form (without bases), which results into a complex number. This scalar product
can be expanded as per Eq.10 once we choose a bases. A useful identity that can be
verified easily using Eq.10 is ⟨ϕ|ψ⟩∗ = ⟨ψ|ϕ⟩. In mathematical jargon there are two LVSs,
conjugate of each other, with element |ψ⟩ having conjugate as ⟨ψ|. When we work with
matrix algebra, it turns out that these two elements are actually Hermitian conjugates of
each other, i.e. (|ψ⟩)† = ⟨ψ|, with the ket being equivalent to a column matrix (or column
vector) and a bra being equivalent to a row matrix.
The wave-function ψ(x) actually gives a complex number at a specific value of x which
basically is the scalar product of the state |ψ⟩ with the basis-state corresponding to x, i.e
|x⟩. So ψ(x) = ⟨x|ψ⟩. One can easily draw an analogy with the LVS of vectors in 3D where

→ −

Vi , i.e. ith component of V , is the scalar product ubi · V . Similarly in p-representation
gψ (p) = ⟨p|ψ⟩. Obviously, gψ (p) and ψ(x) are two different mathematical functions but
they represent the same abstract quantum state. We saw earlier that an abstract operator,
defined over a LVS, gets a concrete form in terms of its matrix elements. The same idea
holds here with a general matrix element of an operator A being ⟨ϕ|A|ψ⟩. It’s insightful
to recall ubj · Aubi for LVS of vectors in 3D. In x-representation, the general matrix element
for operator A would be ⟨x′ |A|x⟩. This is non-trivial since x is a continuous basis, which
we postpone for later; still we can write the matrix  element for the
 Hamiltonian operator
ℏ2 ∂ 2
for a particle in a potential V (x) as ⟨x |H|x⟩ = − 2m ∂x2 + V (x) δ(x′ − x).

We’ll be working with state wave-functions and operators in specific orthonormal bases
discussed in the next section. As we shall see that a wave function |ψ⟩ in a discrete bases
reduces to a set of infinite discrete complex-numbers, ci = ⟨ui |ψ⟩, which we can also
imagine as an infinite column matrix. The Hermitian conjugate of this wave function,
i.e. ⟨ψ|, would then be a row matrix consisting of c∗i . Similarly an operators A in such
discrete bases gets defined by its matrix elements Aij = ⟨ui |A|uj ⟩. Thus an operator can
be easily imagined as a infinite dimensional square matrix with elements Aij and we can
easily define its Hermitian conjugate, A† , consisting of elements A†ij = A∗ji .
This also gives an idea about the algebraic expressions involving bras, kets, opera-
tors and complex numbers. One can, in principle come across four valid types that are
eventually equivalent to

1. a complex number, like c1 ⟨ϕ|A|ψ⟩.

2. a column matrix, like c1 A|ψ⟩.

3. a row matrix, like c1 ⟨ψ|A.

4. a square matrix, like c1 A|ψ⟩⟨ϕ|.

6
The expressions equivalent to product of two kets or two bras (like |ψ⟩∥ϕ⟩ or ⟨ψ|⟨ϕ|) are
unphysical unless one is looking at the direct product states, which is beyond this course.
One often has to convert such complex expressions, such as c1 |ψ1 ⟩⟨ψ2 |A|ψ3 ⟩⟨ψ4 |, into its
Hermitian conjugate, i.e. [c1 |ψ1 ⟩⟨ψ2 |A|ψ3 ⟩⟨ψ4 |]† . For this the prescription is,

1. The order of entities in the product gets reversed and the gets transferred to
individual entities, like operator A becomes A† .

2. The placement of complex numbers (or equivalent expressions) can be arbitrarily


changed as they commute with everything else.

3. For bra and kets we use (|ψ⟩)† = ⟨ψ| or vice versa.

So the expression (c1 |ψ1 ⟩⟨ψ2 |A|ψ3 ⟩⟨ψ4 |)† would simplify to (c∗1 |ψ4 ⟩⟨ψ3 |A† |ψ2 ⟩⟨ψ1 |) or which
is equivalent to (c∗1 ⟨ψ3 |A† |ψ2 ⟩|ψ4 ⟩⟨ψ1 |). A useful exercise at this stage is to analyze the
Hermitian conjugates of A† , cA, AB, A + B, ⟨ψ|A, A|ϕ⟩.
When looking at composite expressions consisting of sums of such terms, one has to
realize that we can only add similar type expressions, i.e. column (row) matrix with a
column (row) matrix, a square matrix with a square matrix and complex number with
another complex number. Thus expressions such as c1 |ψ⟩ + c2 ⟨ϕ| or c1 |ψ⟩ + c2 or c1 |ψ⟩ + A
are meaningless.

2.2 Orthonormal bases for LVS of wave-functions:


The complex LVS of wave-functions is infinite dimensional as it requires infinite number
of basis-vectors (or basis functions) to represent a general wave-function. In fact, there
are two types of basis-sets possible for this LVS, namely, discrete and continuous.
We already have seen examples of two continuous basis-sets, namely x and p. The
continuous bases offer more conceptual difficulty though they are more often used. In
fact, it turns out that the complex-functions corresponding to basis-sets for continuous
bases, such as exp(ipx/ℏ) or δ(x − x0 ) are not normalizable and thus they are not valid
wave-functions. This means that the continuous basis-sets do not belong to the LVS
itself. However, if we take a careful superposition of these basis-functions we can ensure
that the resulting complex functions are valid wave-functions. On the other hand these
basis-functions (like plane-waves) are often used, like in scattering, despite their non-
normalizability. However, one can justify them by putting some envelope (such as a box
function or some other function vanishing at infinity) in real or k-space that such functions
essentially retain their plane wave character but they become normalizable. In this case
one can look at the basis-functions as some sort of limit of a series of wave-functions.
The discrete bases are also encountered in QM and these are easily comprehended.
An example is, the eigen states for a particle in an infinite potential well, i.e. ψn (x) =
p
2/L sin(nπx/L). This set {ψn (x)} gives a discrete set of functions whose superposition
will give any arbitrary wave-function that vanishes at x = 0 & x = L. Another example
is the eigen state wave-functions of a simple harmonic oscillator.
In fact, the complete set of eigen-states corresponding to a given Hamiltonian forms a
basis. However, we may end up getting mixed basis-set in case the Hamiltonian has both
bound states and scattering states. The two examples (box and SHO) that I gave do not

7
have scattering states. It is much easier to discuss the formalism for discrete bases and we
can generalize to continuous bases by replacing the discrete summations by appropriate
integrals. The summations will also span infinite range as this is an infinite-dimensional
LVS. We shall use the Dirac notation and I hope you’ll get more comfortable with it as
we move forward. The Dirac notation is so convenient and popular that I cannot imagine
doing any QM without it.
We start with a discrete but infinite set {ui (x)} of orthonormal basis-functions, i.e.
{u1 (x), u2 (x), u3 (x), u4 (x), ...}. Now these are already written in x-representation, how-
ever we can bring the Dirac notation to represent an abstract basis-state as |ui ⟩ and its
Hermitian conjugate as ⟨ui |. This basically means ui (x) = ⟨x|ui ⟩ and u∗i (x) = ⟨ui |x⟩.
This (or any) basis-set satisfies the following:

1. {|uj ⟩} or {ui (x)} forms a complete set, i.e. any valid state |ψ⟩ with wave-function
ψ(x) can be written as a linear superpositionPof of these bases. P We can write this
in both Dirac notation and otherwise: |ψ⟩ = i ci |ui ⟩ or ψ(x) = i ci ui (x). The ci
values will be same in both ways of writing. Here ci s are complex numbers and the
sum, in general, would run over an infinite range of i.

2. Orthonormality:
Z ∞
⟨ui |uj ⟩ = u∗i (x)uj (x)dx = δij . (14)
−∞

The coefficients ci s represent ui (x)th component inR ψ(x). To find the R c∗i we take the
scalar
R ∗ product
P of |ψ⟩ with
P |uRi ⟩ to get, ⟨ui |ψ⟩ = P ⟨ui |x⟩⟨x|ψ⟩dx = ui (x)ψ(x)dx =

ui (x) j cj uj (x)dx = jP cj ui (x)uj (x)dx P = j cj δij = ci . Thus ci = ⟨ui |ψ⟩ and we
can write ψ(x) = ⟨x|ψ⟩ = i ⟨x|ui ⟩⟨ui |ψ⟩ = i ci ui (x). Effectively, what we have done
here is to introduce a unity operator, i.e.
X
|ui ⟩⟨ui | = 1 (15)

Z ∞
or |x⟩⟨x|dx = 1 (16)
−∞

in between the bra and ket in ⟨x|ψ⟩. The operator in Eq.15 is a sum of operators |ui ⟩⟨ui |,
which is a projection operator that projects out |ui ⟩ component form any given state
|ψ⟩, i.e. |ui ⟩⟨ui |[|ψ⟩] = |ui ⟩⟨ui |ψ⟩ = |ui ⟩ci . One can similarly define a general projection
operator Pϕ = |ϕ⟩⟨ϕ| to project out |ϕ⟩ from any given state.
Also we can work out the scalar product between P abovePψ(x) and another wave-
function ϕ(x) which in {ui (x)} bases is ϕ(x) = i bi ui (x) = Pi ⟨ui |ϕ⟩⟨x|ui ⟩. Using P ∗the
unity operator from Eq.15, the scalar product will give, R ⟨ϕ|ψ⟩ = i ⟨ϕ|ui ⟩⟨u
R i∗|ψ⟩ = i bi ci .
Or
R Pwe ∗can work it out using Eq.16 as ⟨ϕ|ψ⟩ = ⟨ϕ|x⟩⟨x|ψ⟩dx = ϕ (x)ψ(x)dx =

b u (x)c j uj (x)dx which leads to the same final expression after using the orthonor-
ij i i P
mality of {ui (x)}. Also we can see, ⟨ψ|ψ⟩ = i |ci |2 .
The identity operators in Eq.15 & 16 are a statement of closure (or completeness)
for the basis sets {|x⟩} and {|ui ⟩} in the sense that any wave-function can be written as

8
linear superposition of these basis functions.
P There isP another Rrelation that follows from
this completeness: ψ(x) = ⟨x|ψ⟩ = i ⟨x|ui ⟩⟨ui |ψ⟩ = i ui (x) u∗i (x′ )ψ(x′ )dx′ . Here we
have written the
R P scalar product ⟨ui |ψ⟩ using the integral form in x-representation. This
yields ψ(x) = [ i ui (x)u∗i (x′ )] ψ(x′ )dx′ and for this to be true for any arbitrary ψ we
must have
X X
ui (x)u∗i (x′ ) = ⟨x′ |ui ⟩⟨ui |x⟩ = δ(x − x′ ). (17)
i

This is equivalent to Eq.15 (and Eq.16) if one inserts it in ⟨x|x′ ⟩ which is δ(x − x′ ). This
can be thought as the orthonormality condition for continuous x-bases.
As discussed in previous section an operator gets its concrete form in a given bases as
a square matrix, which one can use to find its matrix elements between two general wave-
functions. For instance one can write
P hP ⟨ϕ|A|ψ⟩iby inserting unityP operator of Eq.15 in two
places to get ⟨ϕ∥ [ i |ui ⟩⟨ui |] |A| j |uj ⟩⟨uj | |ψ⟩. This gives ij ⟨ϕ|ui ⟩⟨ui |A|uj ⟩⟨uj |ψ⟩,
P ∗
i.e. ij bi Aij cj i.e. the product of three matrices similar to Eq.5 except for infinite
number of rows and columns in each. Similarly, one can analyze other entities, such as
A|ψ⟩ which will be a column matrix resulting from the product of a square matrix A and
column matrix for |ψ⟩, or ⟨ψ|A which will be a row matrix resulting from the product
of a row matrix corresponding to ⟨ψ| with square matrix A. One P can show that in the
th
former case the i element of the column matrix will be given by j Aij cj while in the
P
latter case the ith element of the row matrix will be j c∗j Aji . It’s is a good exercise to
think of Hermitian conjugates of these two results.

2.3 Bases change using Dirac notation:


If we look at the previous section we basically practiced change of bases between |x⟩-bases
and the discrete |ui ⟩-bases. We can also look at change of bases between two discrete bases,
i.e. from |ui ⟩ to |vi ⟩ and how it transforms the wave functions and operators. In fact one
example is to change from x to p bases and we know that the wave functions transform
according to Eqs.12 & 13. We also discussed in section 1.3 the discrete change of bases but
without using the Dirac notation and for real LVS. The Dirac notation P makes the P bases
change veryPconvenient, P at least notation-wise. We start with |ψ⟩ = i ⟨ui |ψ⟩ = i ci |ui ⟩

and|ψ⟩ = i ⟨v
Pi |ψ⟩ = i ci |vi ⟩. Taking the scalar product of first form Pwith |vj ⟩ we

get ⟨vj |ψ⟩ = c ⟨v |u
i i j i ⟩ and then defining Sji = ⟨v |u
j i ⟩ we get c j = i Sji ci . This

transformation matrix S thus defines the linear relation between ci and ci and the state
vector in |vi ⟩ bases will be given by a column matrix that results when one multiplies S
with the column matrix of |ui ⟩ bases. We can also think of the inverse transformation
and in that case one has to multiply by S −1 with the column matrix in |vi ⟩ bases. For
the two orthonormal bases, S works out to unitary, i.e. SS † = 1, and so S −1 = S † . It’s
a good exercise to prove the unitarity starting from the elements of S, i.e. Sji = ⟨vj |ui ⟩
and using the orthonormality of the two bases.

9
We can also look at how an operator A transforms, as follows. We have,

A′ij = ⟨vi |A|vj ⟩


" # " #
X X
= ⟨vi | |uk ⟩⟨uk | |A| |ul ⟩⟨ul | |vj ⟩
k l
X
= ⟨vi |uk ⟩⟨uk |A|ul ⟩⟨ul |vj ⟩
kl
X
= Sik Akl Slj† .
kl

Thus we can see that the operator in |vi ⟩ bases is given by A′ = SAS † . We can also argue
that the trace and determinant of the operator remains independent of the bases by using
the facts that for a product of two matrices the trace and determinant are independent
of the order of the product.

3 Quantum Operators:
We review some of the properties of operators that are useful for QM. We begin by stating
that the operators associated with physically measurable quantities are always Hermitian
as they are guaranteed to have real eigen values. We recall that a measurement always
yield an eigen value. We have already seen in context of Hamiltonian and Schrödinger
equation what eigen values and eigen states mean. Using Dirac notation we state the
definition again.

3.1 Eigen values and eigen vectors:


A non-trivial (i.e. not null) state |ψ⟩λ is called an eigen state with eigen value λ of an
operator A if

A|ψλ ⟩ = λ|ψλ ⟩. (18)

Theorem-1: If A is Hermitian than all its eigen values λ are real and, conversely, if all
eigen values of A are real than it is Hermitian.

Proof: Given Hermitian A, i.e. A† = A we have A|ψλ ⟩ = λ|ψλ ⟩ = A† |ψλ ⟩. Consider


⟨ψλ |A|ψλ ⟩ = λ⟨ψλ |ψλ ⟩ and ⟨ψλ |A|ψλ ⟩∗ = ⟨ψλ |A† |ψλ ⟩ = λ⟨ψλ |ψλ ⟩ since A† = A. Thus we
get λ⟨ψλ |ψλ ⟩ = λ∗ ⟨ψλ |ψλ ⟩ which implies λ∗ = λ since |ψ⟩λ is not a null state.
Now for the converse, i.e. given all eigen values to be real, we look at A in its eigen
bases. In this bases A will be a diagonal operator with diagonal elements as Eigen values
which are real and thus A is clearly Hermitian in its Eigen bases. Now if we go back to
original (non-eigen) bases using a unitary transformation so that we get SAS † . Now the
task is to prove that if A is Hermitian than SAS † is also Hermitian which we can see
trivially as (SAS † )† = (S † )† A† S † = SAS † .

10
Theorem-2: Two eigen-vectors, corresponding to different eigen values, of a Hermi-
tian operator are orthogonal.

Proof: We have A|ϕ1 ⟩ = λ1 |ϕ1 ⟩ and A|ϕ2 ⟩ = λ2 |ϕ2 ⟩. Now consider ⟨ϕ2 |A|ϕ1 ⟩∗ =
λ∗1 ⟨ϕ2 |ϕ1 ⟩∗ . This is also equal to ⟨ϕ1 |A† |ϕ2 ⟩ which gives λ2 ⟨ϕ1 |ϕ2 ⟩. Subtracting the two, we
get (λ2 − λ1 )⟨ϕ1 |ϕ2 ⟩ = 0. Thus we get the required result that if λ1 ̸= λ2 than ⟨ϕ1 |ϕ2 ⟩ = 0,
i.e. |ϕ1 ⟩ and |ϕ2 ⟩ are orthogonal.

This above theorem-2 will also apply to unitary operators which are also very impor-
tant in QM. The proof will follow the same logic together with the fact that eigen values
of U † and U share have common eigen states but with eigen values that are complex con-
jugate of each other. This is easy to comprehend as a unitary operator U can be written
as U = exp(iλA) with A as a Hermitian operator and λ as a real number. Can we prove
the last sentence?

3.2 Commuting Operators:


Two operators, A & B, that commute with each other, i.e. AB = BA, play an important
role in QM for two reasons:
1. They both can be simultaneous diagonalized, i.e. one can find eigen-states common
to both and thus they will be diagonal in the bases made by these common eigen-
states and,
2. According to Heisenberg equation of motion, if A commutes with Hamiltonian H
than the expectation value of A, over any state, will be time independent and thus
A represents a time invariant observable.
Actually, the operators that commute with H represent certain symmetry of H and this
leads to degeneracy which can be systematically handled by finding common eigen states
of H and the operators that commute with it. For instance invariance of H under space
inversion, i.e. x → −x (or −→r → −− →r ), implies that H commutes with the parity operator
Π. In this case one can always find eigen-states wave functions of H that are also eigen
states of Π. The eigen states of Π are functions that are either even or odd under inver-
sion. In one dimension, there is no degeneracy for bound states and thus we get the eigen
states that are either symmetric or anti-symmetric. Another example that we’ll exten-
sively discuss is the periodic potential where H commutes with translation-by-a operator,
i.e. Ta , with a being the periodicity. Thus we can find the eigen states of H that are also
eigen states of Ta so that these states only change phase when one changes x by a. So
let’s study these commuting operators a bit more systematically.

Theorem-1: If A and B are two commuting operators and if |ψλ ⟩ is an eigen state
of A with eigen value λ than the state B|ψλ ⟩ is also an eigen state of A with the same
eigen value.
Proof: The proof is almost trivial but the implications are very important. Since
AB = BA and A|ψλ ⟩ = λ|ψλ ⟩. Thus we get AB|ψλ ⟩ = BA|ψλ ⟩ = λB|ψλ ⟩ which im-
plies that B|ψλ ⟩ is an eigen state of B. Also, please note that the proof does not require
A & B to be Hermitian.

11
This leads to two possible cases:
1. λ is a non-degenerate eigen value of A, i.e. there is only one state of A with eigen
value λ. In this case B|ψλ ⟩ and |ψλ ⟩ can differ at most by a multiplicative constant,
i.e. B|ψλ ⟩ ∝ |ψλ ⟩ or B|ψλ ⟩ = µ|ψλ ⟩, with µ as a complex number. This implies that
|ψλ ⟩ is also an eigen state of B (with certain eigen value µ).

2. λ is a degenerate eigen value of A, i.e. there are many eigen states, say |ψλi ⟩, of A
with same eigen value λ. In this case we can only assert that B|ψλi ⟩ would be a
linear combination of all such |ψλi ⟩ states and it cannot be an eigen state of A with
different eigen value. In fact B|ψλi ⟩ will be orthogonal to all other eigen states of A
having different eigen values, i.e. ⟨ψµi |B|ψλi ⟩ = 0 with |ψµi ⟩ as an eigen state of A
with eigen value µ (̸= λ).
Theorem-2: If A & B commute, one can construct an orthonormal basis-set of the state
space with the basis vectors being eigen vectors of both A & B. Conversely, If there exists
a basis of eigen vectors common to A and B then A and B commute.

We shall not prove this theorem but illustrate it with an example. Although the proof of
the converse is easy as both A and B will be diagonal operators in the referred common
basis-set of eigen vectors and the diagonal operators always commute. We can easily
construct a simple illustration using A and B in basis |u1 ⟩, |u2 ⟩ & |u3 ⟩ as follows,
         
1 0 0 1 0 0 1 0 0
A = 0 1 0 and B = 0 2 0 . Here, |u1 ⟩ = 0 , |u2 ⟩ = 1 and |u3 ⟩ = 0 .
        
0 0 2 0 0 1 0 0 1

Clearly, AB = BA and both have degeneracies. A has two independent eigen states |u1 ⟩
& |u2 ⟩ with eigen value 1. Thus a general linear combination of the two, i.e. c1 |u1 ⟩+c2 |u2 ⟩
is also an eigen state of A with eigen value 1. The third eigen state of A is |u3 ⟩ with eigen
value 2. For B, b1 |u1 ⟩ + b3 |u3 ⟩ gives the doubly degenerate eigen state with eigen value
1 and |u2 ⟩ is the third eigen state with eigen value 2. As given we already have chosen a
basis encompassing the eigen states of both A & B.
Now one can refer to the three eigen states uniquely by labeling them as (a, b) with a
and b as eigen values wrt A & B. This gives |u1 ⟩ → (1, 1), |u2 ⟩ → (1, 2) and |u3 ⟩ → (2, 1).
Thus B helps in lifting the degeneracy of A completely, or vice-versa, and with systematic
labeling and orthogonal states. The last one definitely holds in case both A and B are
Hermitian or unitary. In general one may need more than two commuting operators to
lift the degeneracy completely.
This was an illustration using Dirac notation and simple 3 dimensional Hilbert space.
We can think of more realistic and physical examples (listed below) in QM where one
takes help of other operators, describing certain symmetry, to lift all the degeneracies and
to systematically label the eigen states.
2
1. Free particle, with Hamiltonian
√ H = p /2m, gives eigen
√ states of H with eigen
value E as ψE (x) = c1 exp(i 2mEx/ℏ) + c2 exp(−i 2mEx/ℏ).. Thus there is
degeneracy of two. Two ways that can be used to lift this degeneracy are: 1) H
commutes with p, 2) H commutes with Π (parity). In the former case we get eigen

12
states common to H and p as exp(ipx/ℏ) with eigen values as p and p2 /2m wrt to
the two operators.
√ In the latter case,
√ we can choose the eigen states common to H
and Π as sin( 2mEx/ℏ) and cos( 2mEx/ℏ) with eigen values as E and ±1 wrt
to the two operators. We note that p and √ Π do not commute√so we cannot find
common eigen-states. More explicitly sin( 2mEx/ℏ) and cos( 2mEx/ℏ) are not
eigen states of p and exp(ipx/ℏ) is not eigen state of Π.

2. Another example that we are familiar with is the hydrogen atom. In this case H
commutes with L2 , i.e. total angular momentum, and Lz , i.e. z-component of
angular momentum. Thus one can lift all the degeneracies of H (excluding spin)
using these two and label the states using eigen values corresponding to the three
operators, i.e. (n, l, m). Here, eigen value wrt H is −n2 E0 , wrt L2 is l(l + 1)ℏ2
and wrt Lz , it is mℏ. Finally L2 and Lz commute with H due to some underlying
symmetries of H.

The set of operators that commute with each other and lift the degeneracy completely
is called complete set of commuting operators (CSCO). This is important in QM as this
gives systematic way to keep track of all the eigen states. The latter helps in keeping
account of states as total number of states with a given energy controls many observ-
able phenomena. We’ll be seeing another problem where these ideas come in handy, i.e.
periodic potential which is important for understanding the physics of electrons in solids.

13
4 Periodic potentials:
In solids the electrons experience the periodic potential due to atomic cores arranged in a
periodic fashion. Different solids have different crystal structure and with different atoms
leading to a variety of properties. To understand these properties we need to first obtain
the energies and wave-functions of the states that the electrons will occupy. It turns out
that these energies consists of continuous bands separated by gaps in between. Finding
this band structure of a given solid is an important problem in solid state physics as this
is the first step towards understanding various properties of a given solid.
We shall start with a general formalism on how to find the possible energy states that
electrons will assume in periodic potentials. The wave-function of these states takes the
form of a periodically modulated plane wave as dictated by Bloch’s theorem and thus
these states are also called Bloch states. Further we shall discuss the periodic boundary
conditions as any given solid has large number of atoms when we look at macroscpic scale
but it’s not mathematically infinite and thus not really periodic in mathematical sense.
This also leads to only certain Bloch states being allowed.

4.1 Bloch’s theorem


Consider a periodic potential V (x) in one dimension with periodicity a so that V (x + a) =
V (x). Thus the time independent Schrödinger equation is
−ℏ2 d2
ψ(x) + V (x)ψ(x) = Eψ(x). (19)
2m dx2
If we change x to x + a in this equation, we get
−ℏ2 d2
ψ(x + a) + V (x + a)ψ(x + a) = Eψ(x + a)
2m dx2
−ℏ2 d2
i.e. ψ(x + a) + V (x)ψ(x + a) = Eψ(x + a)
2m dx2
The last equation implies that ψ(x + a) is also an eigen state of the Hamiltonian H =
−ℏ2 d2
2m dx2
+ V (x), given ψ(x) as an eigen state of H and that too with with the same eigen
energy E. We define an operator Ta , i.e. translation by a so that Ta ψ(x) = ψ(x + a).
The LHS of the two steps that immediately followed Eq.19 can be re-written symbolically
as, Ta Hψ(x) & HTa ψ(x), which are equal since H is invariant under translation by a.
This means H commutes with Ta . This implies from previous chapter that we can find
common eigen states of H and Ta both. The eigen states of Ta have specific structure as
we see further.
We can prove that any arbitrary matrix element of Ta , i.e. ⟨ϕ1 |Ta |ϕ2 ⟩, satisfies
⟨ϕ1 |Ta |ϕ2 ⟩∗ = ⟨ϕ2 |Ta† |ϕ1 ⟩ = ⟨ϕ2 |T−a |ϕ1 ⟩. The proof consists of writing the matrix ele-
ment explicitly in integral form and making a change of variable in the integral from x
to x − a. This implies Ta† = T−a and thus Ta† Ta = 1, i.e. Ta is a unitary operator. One
should think of this with Ta† Ta operating on ψ(x) and giving nothing but ψ(x).
It turns out that any eigen value c of a unitary operator is unimodular, i.e. |c|2 = 1.
The proof starts with the eigen value equation, i.e. Ta |ϕ⟩ = c|ϕ⟩. Now we can take the
norm (or scalar product with itself) of both left and right hand sides to get ⟨ϕ|Ta† Ta |ϕ⟩ =

14
⟨ϕ|c∗ c|ϕ⟩. Using the unitary property of Ta we get ⟨ϕ|ϕ⟩ = |c|2 ⟨ϕ|ϕ⟩ which implies |c|2 = 1
as we are analyzing the non-trivial eigen state |ϕ⟩. The unimodular nature of c implies
that it has a general form c = eiα with α as a real number. It is useful to find alternative
proofs of the statements used here as that leads to more insights.
All this means is that we can always find eigen states of H that are also eigen states
of Ta , i.e. they satisfy Ta ψ(x) = cψ(x) = eiα ψ(x) or

ψα (x + a) = eiα ψα (x). (20)

I have used a subscript α with ψ as it refers to a wave-function having eigen value exp(iα)
wrt operator Ta . Eq.20 is essentially the statement of the Bloch’s theorem but it is usually
stated after imposing the periodic boundary condition which makes the values of α more
explicit as discussed next. We should also keep in mind that both H and Ta may have
degeneracies and at this stage it is not ruled out that combination of H and Ta still leave
some unresolved degeneracies. Use this as food for further thoughts.

4.2 Periodic boundary condition:


We usually look at macroscopic solids which consist of a large number of lattice sites but
however large this is not mathematically infinite and such crystals with finite size are not
mathematically periodic. You must think over this and convince yourself that this is the
case. On the other hand if we look at the physics in the bulk of the crystal and not at
the surface or too close to the surface we expect that in the bulk (far enough from the
surface) the electrons should behave like they are in an infinite crystal. Now keeping the
surface effects aside, can we find a way to make the finite size crystal mathematically
periodic? This will help us in terms of using the above Bloch’s theorem which is valid for
mathematically periodic potentials. We must not forget that the outcome of this theory
should not be used in looking at the surface effects and we have to find other ways to
handle the surface effects.
To make these finite crystals mathematically periodic one uses ”periodic boundary
condition”. Restricting our discussion to 1D, this states that if a given crystal has N
lattice sites we extend it beyond this range, and infinitely, by using the prescription
V (x + N a) = V (x) and correspondingly

ψ(x + N a) = ψ(x). (21)

This can be thought of in 1D as wrapping the lattice into a circle so that (N + 1)th site
coincides with 1st site and everything, including potential and wave-function, repeats when
one completes a full circle. Now this makes the crystal of our interest mathematically
periodic.
Repeating Eq.20 N -times repeatedly, we get ψ(x + N a) = exp(iN α)ψ(x) and then
using in Eq.21 we get exp(iN α) = 1 which gives N α = 2nπ or α = 2nπ/N . Here, n will
range from 0 to N − 1. Beyond this range exp(2nπ/N ) will map to one of the values
for n within 0 to N − 1 for this exp factor. For instance if I take n = N − 1 + n1 I get
exp(2π + 2(n1 − 1)π/N ), i.e a repeat of n − N (=n1 − 1) as e2π is trivially one.
In the end what matters is that we keep n values spanning a range of N different
values. So we might as well use n ∈ [− N2 , N2 − 1] or n ∈ [− N2 + 1, N2 ] and for large N

15
this N2 − 1 or N2 + 1 can be taken as N2 . In solid state physics, for given lattice period
a, one rather uses k = αa = 2πn
Na
in place of α. Here k takes N different values separated
by ∆k = N a = L with L as crystal length. k covers a range k ∈ [− πa , πa − N2πa ] which for
2π 2π

large N can be taken as k ∈ [− πa , πa ]. With k taking place of α we can restate Eq.20 as

ψk (x + a) = eika ψk (x). (22)

This is precisely how the Bloch theorem is stated in most textbooks. We can see that
Bloch wave-function ψk (x) is not periodic except for a special value of k = 0. It turns
out that this the Bloch wave function actually represents a periodically modulated plane
wave, i.e.

ψk (x) = eikx uk (x), (23)

with uk (x) being periodic with lattice period, i.e. uk (x + a) = uk (x). Think over this to
convince yourself on how Eq.23 represents a periodically modulated plane wave. Eq.23
and Eq.22 are equivalent and both are interchangeably used as Bloch’s theorem statement.
It is left as an exercise to prove that Eq.23 indeed satisfies Eq.22. You can also prove the
converse if you feel up to it [Hint: Is e−ikx ψ(x) periodic?].

4.3 Kronig-Penney model

Figure 1: Left: Potential in KP model as an array of finite depth wells with period a.
Right: The KP potential in the δ-function limit.

The band-structure calculations for real systems (even in 1D) are rather cumbersome
and one needs to make many approximations. More than one full semester courses will be
required to learn the state of the art techniques for band structure calculations. However,
to get a flavor I discuss a simple 1D model, called Kronig-Penney (KP) model, which is
also exactly solvable in the δ-function limit. In fact, I’ll only solve the δ-function limit of
this model. The potential for the KP model is shown in Fig.1. The left one shows it as
a periodic array of finite potential wells separated by barriers of height V0 and width b
while the right one shows it in the δ function limit. In the latter case V0 → ∞ and b → 0
such that bV0 = γ stays constant. We can write, in the latter case,
X
V (x) = γ δ(x − na). (24)
n

16
We shall be using periodic boundary conditions with a large N and Bloch’s theorem,
i.e. Eq.22. We do not know the eigen energy (0 < E < ∞) values and we also have to
the corresponding wave-functions. Since the potential vanishes except when x = na we
have the TISE away from these points as −ℏ
2 d2
ψ (x) = Eψk (x). Thus we can write the
2m dx2 k
solution,

for 0 < x < a, ψk (x) = A exp(iλx) + B exp(−iλx) (25)

Here A, B and λ = 2mE/ℏ2 will be dependent on k. Using Eq.23 we can write, ψk (x) =
exp(−ika)ψk (x + a), which can be used to assert,

for − a < x < 0, ψk (x) = exp(−ika)[Aeiλ(x+a) + Be−iλ(x+a) ] (26)

Now this ψk (x) has to satisfy two boundary conditions at x = 0, namely,


1. Continuity, i.e.

ψk (x)|x=0+ = ψk (x)|x=0− (27)

2. Since there is a δ function of strength γ sitting at x = 0, the first derivative of ψk (x)


at this point will have a discontinuity as discussed earlier in the course. This will
be given by

ψk′ |x=0+ − ψk′ |x=0− = 2mγ/ℏ2 (28)

Our goal here is to find A, B and E as a function of k. We use Eq.25 & 26 in Eq.27 to
get

(A + B) = e−ika (Aeiλa + Be−iλa )


or B[1 − e−i(k+λ)a ] = A[e−i(k−λ)a − 1] (29)

and we use Eq.25 & 26 in Eq.28 to get


2mγ
iλ(A − B) − e−ika iλ(Aeiλa − Be−iλa ) = 2 (A + B)
   ℏ 
−i(k+λ)a 2mγ −i(k−λ)a 2mγ
or B −iλ + iλe − 2 = A −iλ + iλe + 2 (30)
ℏ ℏ
Now to eliminate A & B, we cross multiply Eq.29 & Eq.30 to get
2mγ 2mγ
iλ − iλe−i(k+λ)a + − iλe−i(k−λ)a + iλe−2ika − 2 e−i(k−λ)a
ℏ 2 ℏ
−i(k−λ)a 2mγ 2mγ
= −iλ + iλe + 2 + iλe−i(k+λ)a − iλe−2ika − 2 e−i(k+λ)a
ℏ ℏ
−ika iλa −iλa −2ika 2mγ −ika −iλa
or 2iλ − 2iλe [e + e ] + 2iλe − 2 e [e − e−iλa ] = 0

−ika −2ika 2mγ −ika
or 2iλ − 2iλe 2 cos(λa) + 2iλe − 2 e 2i sin(λa) = 0

17
Dividing by 2iλe−ika and introducing P = mγa/ℏ2 , we get

2P
eika + e−ika −2 cos(λa) − sin(λa) = 0
λa
2P
or 2 cos(ka)−2 cos(λa) − sin(λa) = 0
λa
P
or cos(ka) = sin(λa) + cos(λa) (31)
λa
This is the transcendental equation to find valid λ for a given k value. Given this λ(k)
one can also deduce B in terms of k by using Eq.29 (or Eq.30) and A. As usual A will be
dictated by the normalization condition. In order to find λ for given k one has to solve
Eq.31 numerically.

4.3.1 Energy Bands in KP Model


We review a graphical method which is quite insightful for this. Fig.2 shows the plot
P
of λa sin(λa) + cos(λa), i.e. RHS of Eq.31 for P = 5, as a function of λa in units of π.
We know that k will take N different discrete values in interval [−π/a, π/a) separated by
∆k = π/N a and thus cos(ka) will take values between -1 and +1. For instance, for k = 0,
cos(ka) = 1 which is shown as the red horizontal line in Fig.2. This line intersects the
KP-curve at many points. Each intersection point can be projected on the horizontal axis

Figure 2: Plot of the RHS of Eq.31 for P = 5 as a function of λa. The red and blue
horizontal lines represent +1 and -1, respectively, corresponding to ka = 0 & π.

as shown by red dots on x-axis in Fig.2. The λ values corresponding to these dots give

18
energy E = ℏ2 λ2 /2m. Thus for each given k we see that there are many (discrete) energies
possible. Similarly for k = π/a (or −π/a), cos(ka) = −1 and the line corresponding to
this is shown as blue horizontal line, which again intersects KP-curve at many points that
can be projected as blue dots. For any k the horizontal line corresponding to cos(ka)
will lie within these two red and blue lines and thus we see that there are some ranges of
λ that are forbidden as depicted in Fig.3 implying that that certain range (or bands) of
energies are forbidden. The allowed bands of energies are separated by band-gaps and for
a given k there are many (discrete) energies possible with each energy lying in a different
band. We also see that k takes rather continuous values as ∆k = 2π/N a is very small for

Figure 3: Plot of the KP equation with the grey regions showing forbidden bands of λa
responsible for the band gaps in energy.

large N . Thus we get continuous energy bands separated by clear gaps. If we numerically
calculate these energy bands by finding λ values for each k we get E as function of k for
different bands as shown in the left panel of Fig.4 for P = 5. The black lines in this plot
are the bands that are separated by energy gaps from the neighboring ones.
Another point that I want to make is regarding the degeneracy and the Bloch’s theo-
rem. We see that in Eq.31, the cos(ka) term is independent of the sign of k and thus λ as
well as E values are independent of the sign of k implying a degeneracy. This degeneracy
can be traced to the symmetric nature of KP model, i.e. its Hamiltonian commutes with
Π. In detail, we see that ψk and ψ−k wave functions are orthogonal as they are eigen states
of Ta with different eigen values, i.e. e+ika & e−ika . Thus we also see that c1 ψk + c2 ψ−k
is also eigen state of H but not of Ta . Thus c1 ψk + c2 ψ−k does not satisfy the Bloch’s
theorem unless one of the c1 and c2 is zero. In fact the symmetry property of H permits
us to choose eigen states of H that are also eigen states of Π, i.e. if we choose c1 = ±c2 .
Finally, we should understand that Bloch’s theorem, and the choice of Bloch’s states as

19
eigen states of periodic potential, is more of a convention which becomes a rule in case
of no degeneracy. In the end, Bloch states can always be chosen as eigen states of H for
periodic potential.

Figure 4: Left: Energy band diagram E(k) for KP model for P = 5. The red lines mark
the energies corresponding to infinite well potential. Right: plot of the KP equation for
different P values depicting the narrowing of the allowedλ ranges with increasing P .

It is insightful to analyze how this band structure will evolve when one varies P from
zero (i.e. no δ-functions) to infinity. In the later case the potential will be equivalent to
an array of infinite wells. From this we can anticipate the outcome of these two limits.
For P → 0 we should get free particle like energies and for P → ∞ it should give particle-
in-a-box energies. The plot in the right panel of Fig.4 shows the evolution of the RHS of
Eq.31 when P increases from 5 to 40. From this we can see that the bands are expected
to get narrow. In fact the bands in P → ∞ limit collapse on the red lines of the left
plot in Fig.4 which are precisely the energies corresponding to the particle-in-a-box. On
the other hand when P → 0 we see that the energy-gaps will approach zero and we get
an E(k) which is equivalent to the free particle result, i.e. E = ℏ2 k 2 /2m, except that
different portions of the free particle E(k) have been shifted and brought into the k-region
from −π/a to π/a. In this model P is a measure of interaction between the neighboring
potential wells with large P representing small interaction. We see that large interaction
(small P ) we get broad bands and for small interaction (large P ) we get narrow bands.
This is a very general point which gets captured in this simple and exactly solvable model.
More rigorously, if we solve Eq.31 for P → ∞ limit we get λa = nπ and thus E =
n2 π 2 ℏ2
2m
, i.e. the particle-in-a-box energy. However each energy state here has a degeneracy
of N as E is independent of k. In P = 0 limit Eq.31 gives λa = ka + 2nπ and thus
ℏ2
E = 2m (k + 2nπ
a
)2 .

20
4.3.2 Wave-functions in KP Model:
One can also plot some characteristic wave-functions of the KP model to internalize the
modulated plane-wave character of the wave function arising from the Bloch’s theorem.
This requires knowing B as a function of k and A while A, determining the overall scale
of π, can be left out as a normalization constant. From the Bloch wave-function we can

Figure 5: Plots of the Re[ψ], Im[ψ] (red and blue) and |ψ| for the bands and ka as
marked. The last one, with much longer x-range, is for E just above the bottom of the
first band with ka ≃ 0.012π.
list a few general observations as follows:
1. The odd band (1st, 3rd, 5th,..) minimum energy wave functions have k = 0 and

21
max energy wave functions have ka = π.

2. The even band (2nd, 4th, 6th,..) minimum energy wave functions have ka = π and
max energy wave functions have k = 0.

3. For k = 0 the wave functions are periodic with a as ψk=0 (x + a) = ψk=0 (x).

4. For ka = pi the wave functions change sign in every consecutive periodic ψka=π (x +
a) = −ψka=π (x).

5. With increasing energy the average number of nodes per unit length increases. For
lowest energy there are no nodes.

6. The highest energy state of the odd bands correspond to ka = π and those of
the even bands correspond to k = 0 and these states have energy same as that of
particle in an infinite well. These wave functions can also be mapped to infinite well
wave-functions as the wave-functions vanish at the δ-function location, i.e. x = na
leading to no discontinuity in its derivative.

7. Finally we can recognize in the last plot of Fig.5 the modulated plane wave nature as
the wave function, plotted over a large x-range, shows a period-a modulated plane
wave. The plane-wave envelope has a large wavelength given by 2π/k, see Eq.23.
The real and imaginary parts of the wave-function evolve spatially in quadrature.

4.3.3 negative δ-function KP Model


One can also look at the PKP model for a lattice of negative δ-functions, i.e. with overall
potential as V (x) = −γ n δ(x p− na). In this case since the bound state energies will
be negative we redefine λ = −2mE/ℏ2 and P remains as P = mγa/ℏ2 > 0. The
mathematical steps for this are identical and one will get a slight modifications in Eq.31
as the sinusoidal functions of λa will change to the hyperbolic sinusoidal functions. This
equation works out to be cos(ka) = − λa P
sinh(λa) + cosh(λa). In the end we’ll get only
one band of energy. Physically this results from the broadening of the single negative
δ-function bound state. Again it is interesting to analyze the large and small P limits.

4.4 Electrons in solids


Above discussion of KP model gives us a flavor of how electrons behave in a solid. These
E(k) diagrams are referred to as band-structure diagrams resulting from rather complex
calculations incorporating the actual atomic orbitals. We can ignore electron-electron
interaction in many real solids, which, by the way, is not a bad approximation in most
cases. Given the number of electrons, arising from the valence electrons of atoms, we fill
these electrons into these bands respecting Pauli exclusion principle for electrons so that
we put two electrons per energy and per k state. We’ll end up filling certain bands fully
and in several cases we’ll have partially filled bands. In the former case the solid behaves
as an insulator (or semiconductor) as one needs a minimum energy, equal to band gap,
to excite electron. In the later case, i.e. partially filled bands, we get a metal.

22
5 Harmonic Oscillator
Harmonic oscillator is a well know problem in classical mechanics and the quantum har-
monic oscillator is directly useful for understanding the vibrational spectra of molecules
and the lattice vibrations (or phonons) in solids. Indirectly, it is found useful in cases such
as free electrons in uniform magnetic field giving rise to Landau levels. The latter plays
a critical role in quantum Hall effect. We shall discuss the eigen energies and associated
wave-functions of the 1D harmonic oscillator in detail. There are two ways to do this,
one is the brute force solution of the differential equation, i.e. TISE, and the other em-
ploys raising and lowering operators in an elegant way to come to the eigen energies and
wave-functions. The latter is a technique which is useful in many places, such as angular
momentum and many body physics. It is easier to work out some of the quantities by
using the second method.
Let’s begin by stating the Hamiltonian for the SHO as
p2 1 p2 1
H= + kx2 = + mω 2 x2 (32)
2m 2 2m 2
p
t’s useful to non-dimensionalize x by defining ξ = x/x0 with x0 = ℏ/mω which gives,

p2 1 p2 1 ℏ
H= + mω 2 ξ 2 x20 = + mω 2 ξ 2
2m  2  2m 2 mω
ℏω p 2
or H = + ξ2 (33)
2 mℏω
I leave it as an exercise to convince yourself that x0 has dimensions of length.

5.1 First Method: solving the TISE for eigen states


In this equation we can also use the quantum operator p = −iℏ dx
d
or rather after using
ℏ d
x = x0 ξ we substitute p = −i x0 dξ to get
   
ℏω ℏ2 d2 ℏω ℏ2 mω d2
H= − 2
+ξ = − +ξ 2
2 mℏωx20 dx2 2 mℏω ℏ dx2
 
ℏω d2
or H = − 2 +ξ 2
(34)
2 dξ
In order to solve for eigen states we need to write the TISE, i.e. Hψ = Eψ. Here we also
non-dimensionalize E using E = ℏω 2
ϵ and then use H from Eq.34 to get,

d2
− ψ(ξ) + ξ 2 ψ(ξ) = ϵψ(ξ). (35)
dξ 2
Please note that ψ here is taken as a function of scaled x, i.e. ξ, which can always be
converted back to a function of x. The simple looks of this differential equation are
deceiving as it is not so easy to solve. Our objective here is to find physically acceptable
ψ(ξ), i.e. square integrable, and corresponding ϵ values. We want ψ(ξ) to approach
zero for ξ → ±∞. In this limit Eq.35 can be written up to the leading order in ξ as

23
d2
dξ 2
ψ(ξ) = ξ 2 ψ(ξ), which for large ξ will admit solutions of form ξ n exp(±ξ 2 /2). You can
try this but make sure that you keep only the leading order terms in ξ. This provides us
a hint that a valid wave function should have a form ψ(ξ) ∼ u(ξ) exp(−ξ 2 /2), at large ξ
with u(ξ) as a polynomial function. We note that exp(+ξ 2 /2) blows up at ∞. Thus we
substitute ψ(ξ) = u(ξ) exp(−ξ 2 /2) in Eq.35 to find a differential equation for u(ξ). We
−ξ 2 /2)
d2
] = u′′ e−ξ /2 − 2u′ ξe−ξ /2 + uξ 2 e−ξ /2 − ue−ξ /2 in Eq.35 to get
2 2 2 2
use dξ 2 [ue

−u′′ e−ξ /2 + 2u′ ξe−ξ /2 − uξ 2 e−ξ /2 + ue−ξ /2 + uξ 2 e−ξ = ϵue−ξ


2 2 2 2 2 /2 2 /2

or − u′′ +2u′ ξ − uξ 2 + u + uξ 2 = ϵu
or − u′′ +2u′ ξ + u = ϵu
or u′′ − 2u′ ξ + (ϵ − 1)u = 0 (36)

This equation turns out to be Hermite equation which gives polynomial solutions, i.e. a
sum of powers of ξ with the maximum power of ξ as a finite positive integer. But the
polynomial solution exist only if the coefficient of u in the last term on left of Eq.35 is an
even integer, i.e. ϵ − 1 = 2n. This also ensures that overall ψ(ξ) = u(ξ)e−ξ /2 goes to zero
2

for ξ → ±∞. Thus we see that the energies are quantized as ϵ = 2n + 1 or


 
ℏω 1
En = (2n + 1) = n + ℏω. (37)
2 2

The solutions to Eq.35 are called Hermite polynomials, i.e. Hn (ξ), which for a given n is
2 dn −ξ 2
given by Hn (ξ) = (−1)n eξ dξ ne . You can use this to work out a few, i.e. H0 (ξ) = 1,
H1 (ξ) = 2ξ, H2 (ξ) = 4ξ − 2. Eventually, the overall eigen-state wave-function (normal-
2

ized) turns out to be,


r r   mω 
mω mω
ψn (x) = √ n Hn x exp − x2 . (38)
ℏ π2 n! ℏ 2ℏ

We can plot a few wave-functions as shown shown in Fig.6. A few remarks:

Figure 6: Plots of the SHO wave-functions for different n values. Note the xaxis scale
here is x0 and in the right figure the scale is longer.

24
1. These ψn (x) are orthogonal to each other and form an orthonormal basis-set (i.e.
a complete set) for LVS of complex wave-functions. It is useful to iterate that the
most general wave-function for SHO need not be an energy eigen state but it can
always be written as a linear combination of these eigen states. In the latter case
the wave-function can significantly evolve with time.

2. The lowest energy, i.e. E0 = ℏω/2, wave function ψ0 (x) is a Gaussian, which
happens to be the minimum uncertainty wave-function.

3. The nth state wave function has n nodes.

The spread of the wave-function increases with n. In fact for nth state ∆x ∼
4. p
(n + 1)x0 .

5. For large n the probability near classical turning points increases, which is can be
expected as the SHO spends more time near these points.

5.2 Second method: eigen states using special operators



We begin by non-dimensionalizing p in Eq.33 by defining η = p/p0 with p0 = mℏω
which gives
ℏω 2
H= (η + ξ 2 ). (39)
2
p √
We note that x0 p0 = ℏ/mω mℏω = ℏ. Since ξ and η are related to x and p we can
find their commutator as [ξ, η] = x01p0 [x, p] = iℏℏ = i. Now we define two operators,

1 1
a+ = √ (η + iξ) and a− = √ (η − iξ). (40)
2 2
Please note and convince yourself, that these operators, being Hermitian conjugate of
each other, are not Hermitian themselves. You can workout the commutator of these
which works out to be

[a+ , a− ] = −1 (41)

and thus [a− , a+ ] = 1. Also we can workout the product of the two, i.e. a+ a− = 12 (η +
iξ)(η − iξ). We have to remember that the order of ξ and η is important to keep track of
as they don’t commute. You’ll finally get a+ a− = 21 (η 2 + ξ 2 − 1). Thus we can write the
Hamiltonian in Eq.39 as
 
ℏω 1
H= (2a+ a− + 1) = ℏω a+ a− + . (42)
2 2
Before beginning the logical solution let’s also find the commutator [a+ a− , a+ ]. We can
use identity [AB, C] = A[B, C]+[A, C]B and Eq.41 to show [a+ a− , a+ ] = a+ and similarly
[a+ a− , a− ] = −a− . Using this we can show, [H, a+ ] = ℏωa+ and [H, a− ] = −ℏωa− .
Now that we have done all the relevant arithmetics we can start the logical arguments
as follows. Suppose ψ(x) is an eigen state of H with energy E and ask if a+ ψ(x) will

25
also be an eigen state of H and if yes than what is the eigen energy? You should ponder
over this for a minute or two and see if you can get the answer. Anyhow from above we
see [H, a+ ]ψ(x) = ℏωa+ ψ(x) and after expanding the commutator we get Ha+ ψ(x) =
a+ Hψ(x) + ℏωa+ ψ(x). Now using Hψ(x) = Eψ(x) we get Ha+ ψ(x) = a+ Eψ(x) +
ℏωa+ ψ(x) = (E + ℏω)a+ ψ(x). This implies that a+ ψ(x) is also an eigen state of H but
with eigen energy E + ℏω. An exactly identical chain of arguments replacing a+ by a−
leads to the conclusion that that a− ψ(x) is also an eigen state of H but with eigen energy
E − ℏω. Thus we see that a− lowers the energy of an eigen state by ℏω and a+ raises the
energy be sale amount. Therefore a− is called lowering operator and a+ is called raising
operator. These are also called ladder operators as they lead to different eigen states like
the steps of a ladder.
Now if we apply a− repeatedly on ψ(x) we’ll get eigen states with energies E − ℏω,
E − 2ℏω, E − 3ℏω, E − 4ℏω, .... But we remember that since the minimum of the given
potential is at zero we cannot have bound states with negative energies. The only logical
way to get out of this reduction in eigen energy to below zero value is to end up with a
null state at some point, i.e. there exists an energy eigen state ϕ(x) for which a− ϕ(x) = 0.
Now any further application of a− on a− ϕ(x) will only give a null state and we cannot go
below certain energy. This special state ϕ(x) is an eigen state of a− as well as a+ a− with
zero eigen values wrt both. Using Eq.42 we see that Hϕ(x) = ℏω 2
ϕ(x). i.e. this ϕ(x) is an
eigen state of H with energy ℏω 2
and since there is no lower energy state this represents
the ground state energy. To get the ground state wave-function we can solve explicitly
a− ϕ(x) = 0 or (η + iξ)ϕ(x) = 0 or ( pp0 + i xx0 )ϕ(x) = 0. This gives,
 r 
iℏ d mω dϕ mω
−√ −i ϕ(x) =0 or =− xϕ, which leads to
mωℏ dx ℏ dx ℏ
mω 2
ϕ(x) = c0 exp(− x ). (43)
2ℏ
This ϕ(x) is nothing but ψ0 (x) found earlier. We can find c0 from the normalization
condition. We can easily find the higher energy wave functions by operating by a+ on
this wave-function repeatedly, i.e. a+ ψ0 (x) will give the state with energy 32 ℏω and thus
a+ ψ0 (x) = c1 ψ1 (x) and eventually (a+ )n ψ0 (x) = cn ψn (x). Here we need to ensure that
cn is such that ψn is normalized given that ψ0 is normalized. We also see that the
resulting states ψn (x) are eigen states of H with eigen energy same as that given by
Eq.37. Obviously ψn (x) are also eigen states of operator N = a+ a− with eigen values n
and thus it is also called the number operator. Working out of the normalization factor
p an exercise which involves evaluating the norm of a+ |ψn ⟩. It turns out that
cn is left as
a+ |ψn ⟩ = (n + 1)|ψn+1 ⟩. Here, one should remember that a+ pa− |ψn ⟩ = n|ψn ⟩ and using
Eq.41 a− a+ = a+ a− + 1 . One can also show that a− |ψn ⟩ = (n)|ψn−1 ⟩. Finally, using
these recursively, the nth eigen state works out as ψn (x) = √1n! (a+ )n ψ0 (x) with energy
En = (n + 21 )ℏ]ω.
The second method is rather elegant and it also turns out to be easier to evaluate
some of the quantities using a− & a+ operators. One can for instance write physically
relevant x and p operators, and their powers, in terms of a− & a+ . This leads to rather
easy evaluation of x and p uncertainties.

26
5.3 Harmonic oscillator in higher dimensions
The solution of 1D SHO can easily be used in discussing the 2D and 3D SHO. In 2D,
1
the isotropic SHO will be given by Hamiltonian, H2D = 2m (p2x + p2y ) + 12 mω 2 (x2 + y 2 )
1 1
and in 3D it’ll be H3D = 2m (p2x + p2y + p2z ) + 2 mω 2 (x2 + y 2 + z 2 ). We can see from the
form of H that since there are no coupled terms in potential of x, y & z we can separate
the TISE into separate differential equations of x, y & z. Thus for 3D we can write the
overall wave-function as a product of three independent wave-functions, i.e. ψ(x, y, z) =
ℏ2 d2 ψx (x)
ψx (x)ψy (y)ψz (z) with the three satisfying equations, − 2m dx2
+ 12 mω 2 x2 = Ex ψx (x),
ℏ2 d2 ψy (y) ℏ2 d2 ψz (z)
− 2m dy 2
+ 12 mω 2 y 2 = Ey ψy (y) and − 2m dz 2
+ 21 mω 2 z 2 = Ez ψz (z). The total energy
will be given by E = Ex + Ey + Ez .
There will be quantization
 of energy wrt to all the degrees of freedom leading  to
Ex,y,z = nx,y,z + 2 ℏω and thus the total energy will be E = nx + ny + nz + 2 ℏω. It
1 3

is interesting to analyze the degeneracies of different energy states in this case. So for
example the ground state will have energy 23 ℏω and it’s non-degenerate. The first excited
state will have energy 52 ℏω but it will be 3-fold degenerate as any of nx,y,z can assume a
value one. I leave  it to you to analyze the general state’s degeneracy, i.e. the one having
energy N + 2 ℏω. A hint for this is: find the number of ways of arranging n balls and
3

3 bars in a row. One can also generalize the above discussion for 2D SHO.

27

You might also like