ph217 Vip
ph217 Vip
∑|𝑐𝑛 (𝑡)|2 = 1
𝑛
In quantum mechanics, when we make a measurement on a small system with a large
measuring apparatus, of some quantity such as energy, we find the following behavior,
which is sometimes elevated to a postulate or hypothesis in quantum mechanics:
On measurement, the system collapses into an eigenstate of the quantity being measured,
with probability
𝑃𝑛 = |𝑐𝑛 |2
where 𝑐𝑛 is the expansion coefficient in the (orthonormal) eigenfunctions of the quantity
being measured.
Suppose now that we measure the energy of our system in such an experiment. We could
repeat the experiment many times and get a statistical distribution of results. Given the
probabilities, we would find in the usual way that the average value of energy 𝐸 that we
would measure would be
〈𝐸〉 = ∑ 𝐸𝑛 𝑃𝑛 = ∑ 𝐸𝑛 |𝑐𝑛 |2
𝑛 𝑛
where we are using the notation E to denote the average value of 𝐸, a quantity we call the
“expectation value of 〈𝐸〉 in quantum mechanics, the 𝐸𝑛 are the energy eigenvalues.)
For example, for the coherent state discussed above with parameter 𝑁, we have
∞
𝑁 𝑛 exp(−𝑁)
〈𝐸〉 = ∑ 𝐸𝑛
𝑛!
𝑁=0
∞
𝑁 𝑛 exp(−𝑁) 1
= ℏ𝜔 [∑ 𝑛 ] + ℏ𝜔
𝑛! 2
𝑁=0
1
= (𝑁 + ) ℏ𝜔
2
We can show that having an energy ≈ 𝑁ℏ𝜔 for the large 𝑁 implicit in a classical situation
corresponds very well to our notions of energy, frequency and oscillation amplitude in a
classical oscillator. Note that N is not restricted to being an integer – it can take on any real
value. Quite generally, the expectation value of the energy or any other quantum-mechanically
measurable quantity is not restricted to being one of a discrete set of values – it can take on any
real value within the physically possible range of the quantity in question.
We can now show an important but simple relation between the Hamiltonian operator, the
wavefunction, and the expectation value of the energy. Consider the integral
̂ Ψ(𝑟, 𝑡)𝑑3 𝑟
𝐼 = ∫ 𝛹 ∗ (𝑟, 𝑡) 𝐻
Where Ψ(𝑟, 𝑡) is the wavefunction of some system of interest. We can expand this
wavefunction in energy eigenstates, as in Ψ(𝑟, 𝑡) = ∑𝑛 𝑐𝑛 (𝑡)𝜓𝑛 (𝑟). We know that, with 𝜓𝑛 (𝑟)
as the energy eigenstates (of the time-independent Schrödinger equation)
ℏ2 2 ℏ2 2
𝐻̂ 𝛹(𝑟, 𝑡) = [− ∇ + 𝑉(𝑟, 𝑡)] 𝛹(𝑟, 𝑡) = [− 𝛻 + 𝑉(𝑟, 𝑡)] ∑ 𝑐𝑛 (𝑡)𝜓𝑛 (𝑟)
2𝑚 2𝑚
𝑛
= ∑ 𝑐𝑛 (𝑡)𝐸𝑛 𝜓𝑛 (𝑟)
𝑛
And also
+∞
̂ Ψ(𝑟, 𝑡)𝑑3 𝑟 = ∫
∫ 𝛹 ∗ (𝑟, 𝑡) 𝐻 ∗ (𝑡)𝜓 ∗ (𝑟)
[∑ 𝑐𝑚 𝑚 ] [∑ 𝑐𝑛 (𝑡)𝐸𝑛 𝜓𝑛 (𝑟)] 𝑑 3 𝑟
−∞ 𝑚 𝑛
Given the orthonormality of the 𝜓𝑛 (𝑟) , we have
̂ Ψ(𝑟, 𝑡)𝑑3 𝑟 = ∑ 𝐸𝑛 |𝑐𝑛 |2
∫ 𝛹 ∗ (𝑟, 𝑡) 𝐻
𝑛
The expectation value of the energy is
̂ 𝛹(𝑟, 𝑡)𝑑 3 𝑟
〈𝐸〉 = ∫ 𝛹 ∗ (𝑟, 𝑡) 𝐻
The expectation value of the position is:
〈𝑥〉 = ∫ 𝑥 |𝛹(𝑥, 𝑡)|2 𝑑𝑥
+∞ +∞
+∞
𝜕 𝛹 ∗ (𝑥, 𝑡) 𝜕𝛹(𝑥, 𝑡)
= 𝑚 ∫ 𝑑𝑥 ( 𝑥𝛹(𝑥, 𝑡) + 𝛹 ∗ (𝑥, 𝑡) 𝑥 )
𝜕𝑡 𝜕𝑡
−∞
𝑑〈𝑥〉
is the velocity of the expectation value of 𝑥, not the velocity of the particle.
𝑑𝑡
𝑑𝑥
Since classically 𝑝 = 𝑚𝑣 = 𝑚 𝑑𝑡 .
Note that there is no 𝑑𝑥⁄𝑑𝑡 under the integral sign. The only quantity that varies with time is
𝛹(𝑥, 𝑡), and it is this variation that gives rise to a change in 〈𝑥〉 with time. Use the Schrodinger
equation and its complex conjugate to evaluate the above and we have
+∞
ℏ 𝜕 2𝛹∗ ∗
𝜕 2𝛹
〈𝑝〉 = ∫ 𝑑𝑥 ( 2 𝑥𝛹 − 𝛹 𝑥 )
2𝑖 𝜕𝑥 𝜕𝑥 2
−∞
Now
𝜕2 𝛹∗ 𝜕 𝜕𝛹 ∗ 𝜕𝛹 ∗ 𝜕𝛹 ∗ 𝜕𝛹
𝑥𝛹 = 𝜕𝑥 [ 𝜕𝑥 𝑥𝛹] − 𝛹- 𝜕𝑥 𝑥 𝜕𝑥
𝜕𝑥 2 𝜕𝑥
𝜕 𝜕𝛹 ∗ 𝜕 𝜕𝛹 𝜕 𝜕𝛹 𝜕𝛹 𝜕2 𝛹
= 𝜕𝑥 [ 𝜕𝑥 𝑥𝛹] − 𝜕𝑥 (𝛹 ∗ 𝛹)-𝛹 ∗ − 𝜕𝑥 [𝛹 ∗ 𝑥 ] + 𝛹∗ + 𝛹 ∗𝑥
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥 2
Because the wave functions vanish at infinity, the first term does not contribute, and the
integral gives
+∞
ℏ 𝜕
〈𝑝〉 = ∫ 𝑑𝑥 𝛹 ∗ (𝑥, 𝑡) ( ) 𝛹(𝑥, 𝑡)
𝑖 𝜕𝑥
−∞
+∞
As the position expectation was represented by 〈𝑥〉 = ∫−∞ 𝛹 ∗ 𝑥 𝛹𝑑𝑥
Remember:
∞
2 1 𝜋 1⁄2
∫ 𝑒 −𝑎𝑥 𝑑𝑥 = ( )
0 2 𝑎
∞
2 −𝑎𝑥 2
1 𝜋 1⁄2
∫ 𝑥 𝑒 𝑑𝑥 = ( )
0 4𝑎 𝑎
ℏ 𝜕 2 𝜕 2𝛹
〈𝑝2 〉 = ∫ 𝛹 ∗ ( ) 𝛹𝑑𝑥 = −ℏ2 ∫ 𝛹 ∗ 2 𝑑𝑥
𝑖 𝜕𝑥 𝜕𝑥
2𝑎𝑚 2𝑎𝑚𝑥 2
= −ℏ2 ∫ 𝛹 ∗ [− (1 − ) 𝛹] 𝑑𝑥 = 2𝑎𝑚ℏ{∫|Ψ|2 𝑑𝑥
ℏ ℏ
2𝑎𝑚
− ∫ 𝑥 2 |Ψ|2 𝑑𝑥}
ℏ
2𝑎𝑚 2 2𝑎𝑚 ℏ 1
= 2𝑎𝑚ℏ (1 − 〈𝑥 〉) = 2𝑎𝑚ℏ (1 − ) = 2𝑎𝑚ℏ ( ) = 𝑎𝑚ℏ
ℏ ℏ 4𝑎𝑚 2
ℏ ℏ
(c) 𝜎𝑥2 = 〈𝑥 2 〉 − 〈𝑥 〉2 = 4𝑎𝑚 ⟹ 𝜎𝑥 = √4𝑎𝑚
𝜎𝑝2 = 〈𝑝2 〉 − 〈𝑝 〉2 = 𝑎𝑚ℏ ⟹ 𝜎𝑝 = √𝑎𝑚ℏ
ℏ ℏ
𝜎𝑥 𝜎𝑝 = √4𝑎𝑚 √𝑎𝑚ℏ = 2
𝜕𝛹(𝑥, 𝑡) ̂
𝑖𝐻
= − 𝛹(𝑥, 𝑡)
𝜕𝑡 ℏ
̂ does not depend explicitly on time (i.e., the potential 𝑉(𝑥)is constant)
And presuming 𝐻
could we somehow legally write
̂ (𝑡1 − 𝑡𝑜 )
𝑖𝐻
𝛹(𝑥, 𝑡1 ) = 𝑒𝑥𝑝 {= − 𝛹(𝑥, 𝑡𝑜 )}
ℏ
̂ here was replaced by a constant number
Certainly, if the Hamiltonian operator 𝐻
we could perform such an integration of
𝜕𝛹(𝑥, 𝑡) ̂
𝑖𝐻
= − 𝛹(𝑥, 𝑡)
𝜕𝑡 ℏ
to get
̂ (𝑡1 − 𝑡𝑜 )
𝑖𝐻
𝛹(𝑥, 𝑡1 ) = 𝑒𝑥𝑝 {− 𝛹(𝑥, 𝑡𝑜 )}
ℏ
If, with some careful definition, it was legal to do this then we would have an operator that
gives us the state at time 𝑡1 directly from that at time 𝑡𝑜 .
To think about this “legality” ,first we note that, because is a linear operator for any number a
̂ [𝑎𝛹(𝑟, 𝑡)] = 𝑎𝐻
𝐻 ̂ 𝛹(𝑟, 𝑡)
Since this works for any function, 𝛹(𝑟, 𝑡) we can write as a shorthand
̂ 𝑎 = 𝑎𝐻
𝐻 ̂
𝜓 (𝑟) = ∑ 𝑎𝑛 𝜓𝑛 (𝑟)
𝑛
Then we know multiplying by the complex exponential factors for the time-evolution of each
basis function
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝛹(𝑟, 𝑡1 ) = ∑ 𝑎𝑛 𝑒𝑥𝑝 [− ] 𝜓𝑛 (𝑟)
ℏ
𝑛
𝑥2 𝑥3
noting that exp(𝑥) = 1 + 𝑥 + 2𝑖 + 3𝑖 + ⋯
̂ (𝑡1 − 𝑡𝑜 )
𝑖𝐻 1 ̂ (𝑡1 − 𝑡𝑜 ) 2
𝑖𝐻
𝛹(𝑟, 𝑡1 ) = ∑ 𝑎𝑛 [1 + [− ] + [− ] + ⋯ ] 𝜓𝑛 (𝑟)
ℏ 2𝑖 ℏ
𝑛
̂ commute with scalar quantities (numbers) we can
because the operator and all its powers 𝐻
rewrite
2
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 1 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝛹(𝑟, 𝑡1 ) = [1 + [− ] + [− ] + ⋯ ] ∑ 𝑎𝑛 𝜓𝑛 (𝑟)
ℏ 2𝑖 ℏ
𝑛
2
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 1 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝛹(𝑟, 𝑡1 ) = [1 + [− ] + [− ] + ⋯ ] 𝛹(𝑟, 𝑡𝑜 )
ℏ 2𝑖 ℏ
So, provided we define the exponential of the operator in terms of a power series, i.e.,
2
𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 ) 1 𝑖𝐸𝑛 (𝑡1 − 𝑡𝑜 )
𝑒𝑥𝑝 [− ] = [1 + [− ] + [− ] +⋯]
ℏ ℏ 2𝑖 ℏ
Hence, we have established that there is a well-defined operator that given the quantum
mechanical wavefunction or “state” at time 𝑡𝑜 will tell us what the state is at a time 𝑡1
• Schwarz inequality
For any two states|𝜓⟩and |𝜙⟩of the Hilbert space, we can show that
|⟨𝜓|𝜙⟩|2 ≤ ⟨𝜓|𝜓⟩⟨𝜙|𝜙⟩
If|𝜓⟩and |𝜙⟩are linearly dependent (i.e., proportional: |𝜓⟩ = 𝛼|𝜙⟩, where 𝛼 is a scalar), this
relation becomes an equality. The Schwarz inequality (12) is analogous to the following
relation of the real Euclidean space;
2 2 2
|𝐴⃗. 𝐵
⃗⃗ | ≤ |𝐴⃗| |𝐵
⃗⃗ | .
• Triangle inequality
√⟨𝜓 + 𝜙|𝜓 + 𝜙⟩ ≤ √⟨𝜓|𝜓⟩ + √⟨𝜙|𝜙⟩
If|𝜓⟩and |𝜙⟩are linearly dependent, |𝜓⟩ = 𝛼|𝜙⟩, and if the proportionality scalar 𝛼 is real and
positive, the triangle inequality becomes an equality. The counterpart of this inequality in
Euclidean space is given by |𝐴⃗ + 𝐵⃗⃗ | ≤ |𝐴⃗| + |𝐵
⃗⃗ |
• Orthogonal states
Two kets, |𝜓⟩and |𝜙⟩, are said to be orthogonal if they have a vanishing scalar product:
⟨𝜓|𝜙⟩ = 0
• Orthonormal states
Two kets, |𝜓⟩and |𝜙⟩, are said to be orthonormal if they are orthogonal and if each one of them
has a unit norm:
⟨𝜓|𝜙⟩ = 0, ⟨𝜓|𝜓⟩ = 1, ⟨𝜙|𝜙⟩ = 1
• Forbidden quantities
If |𝜓⟩and |𝜙⟩, belong to the same vector (Hilbert) space, products of the type |𝜓⟩|𝜙⟩ and
⟨𝜓|⟨𝜙|are forbidden. They are nonsensical, since |𝜓⟩|𝜙⟩ and ⟨𝜓|⟨𝜙| are neither kets nor bras
(an explicit illustration of this will be carried out in the example below and later on when we
discuss the representation in a discrete basis). If |𝜓⟩ and |𝜙⟩ belong, however, to different
vector spaces (e.g., |𝜓⟩ belongs to a spin space and |𝜙⟩to an orbital angular momentum space),
then the product |𝜓⟩|𝜙⟩written as|𝜓⟩⨂|𝜙⟩, represents a tensor product of|𝜓⟩ and|𝜙⟩. Only in
these typical cases are such products meaningful.
Example1: (Note: We will see later in this chapter that kets are represented by column matrices
and bras by row matrices; this example is offered earlier than it should because we need to
show some concrete illustrations of the formalism.) Consider the following two kets:
−3𝑖 2
|𝜓⟩ = (2 + 𝑖 ), |𝜙⟩ = ( −𝑖 )
4 2 − 3𝑖
(a) Find the bra ⟨𝜙|.
(b) Evaluate the scalar product ⟨𝜙|𝜓⟩.
(c) Examine why the products |𝜓⟩ |𝜙⟩and ⟨𝜙|⟨𝜓|do not make sense.
Solution:
(a) As will be explained later when we introduce the Hermitian adjoint of kets and bras,
we want to mention that the bra⟨𝜙| can be obtained by simply taking the complex
conjugate of the transpose of the ket |𝜙⟩:
⟨𝜙| = (2 𝑖 2 + 3𝑖)
(b) The product ⟨𝜙|𝜓⟩ can be calculated as follows:
−3𝑖
⟨𝜙|𝜓⟩ = (2 𝑖 2 + 3𝑖) (2 + 𝑖 )
4
= 2(−3𝑖) + 𝑖(2 + 𝑖) + 4(2 + 3𝑖) = 7 + 8𝑖
(c) First, the product |𝜓⟩ |𝜙⟩ cannot be performed because, from linear algebra, the product
of two column matrices cannot be performed. Similarly, since two row matrices cannot
be multiplied, the product ⟨𝜙|⟨𝜓| is meaningless.
With |ψ𝑛 ⟩ and E𝑛 as the energy eigenfunctions and eigenvalues of the time independent
equation
̂𝑜 |ψ𝑛 ⟩ = E𝑛 |ψ𝑛 ⟩
𝐻
we expand the solution of the time-dependent Schrödinger equation |Ψ⟩ as:
̂𝑜 +𝐻
∑(𝑖ℏ𝑎̇ 𝑛 + 𝑎𝑛 E𝑛 )𝑒𝑥𝑝(−𝑖 E𝑛 𝑡⁄ℏ)|ψ𝑛 ⟩ = ∑ 𝑎𝑛 (𝐻 ̂𝑝 (𝑡)) 𝑒𝑥𝑝(−𝑖 E𝑛 𝑡⁄ℏ)|ψ𝑛 ⟩
𝑛 𝑛
𝜕𝑎𝑛
where 𝑎̇ 𝑛 ≡ 𝜕𝑡
̂𝑝 (𝑡)|ψ𝑛 ⟩
𝑖ℏ𝑎̇ 𝑞 (𝑡)𝑒𝑥𝑝(− iE𝑞 𝑡⁄ℏ) = ∑ 𝑎𝑛 (𝑡)𝑒𝑥𝑝(−𝑖 E𝑛 𝑡⁄ℏ) ⟨𝜓𝑞 |𝐻
𝑛
We have made no approximations in going from the time-dependent Schrödinger equation to
the above equation, these are entirely equivalent equations.
Now we consider a perturbation series. We introduce the expansion parameter 𝛾 just as before,
̂𝑝
now writing our perturbation as 𝛾𝐻
As before, we can set this parameter to a value of 1 at the end. We presume that we can express
the expansion coefficients n a as a power series:
(0) (1) (2)
𝑎𝑛 = 𝑎𝑛 + 𝛾𝑎𝑛 + 𝛾 2 𝑎𝑛 + ⋯
and we substitute this expansion into the derived equation above:
̂𝑝 (𝑡)|ψ𝑛 ⟩
𝑖ℏ𝑎̇ 𝑞 (𝑡)𝑒𝑥𝑝(−𝑖 E𝑞 𝑡⁄ℏ) = ∑ 𝑎𝑛 (𝑡)𝑒𝑥𝑝(− iE𝑛 𝑡⁄ℏ) ⟨𝜓𝑞 |𝐻
𝑛
Equating powers of 𝛾 , we obtain for the zero order term
(0)
𝑎̇ 𝑞 (𝑡)=0
The zero order solution simply corresponds to the unperturbed solution, and hence there is no
change in the expansion coefficients in time. Repeating the relevant equations
(0) (1) (2)
𝑎𝑛 = 𝑎𝑛 + 𝛾𝑎𝑛 + 𝛾 2 𝑎𝑛 + ⋯
̂𝑝 (𝑡)|ψ𝑛 ⟩
𝑖ℏ𝑎̇ 𝑞 (𝑡)𝑒𝑥𝑝(−𝑖 E𝑞 𝑡⁄ℏ) = ∑ 𝑎𝑛 (𝑡)𝑒𝑥𝑝(− iE𝑛 𝑡⁄ℏ) ⟨𝜓𝑞 |𝐻
𝑛
For the first order term, we have
(1) 1 (0)
𝑎̇ 𝑞 (𝑡) = ̂𝑝 (𝑡)|ψ𝑛 ⟩
∑ 𝑎𝑛 𝑒𝑥𝑝(𝑖𝜓𝑞𝑛 𝑡) ⟨𝜓𝑞 |𝐻
𝑖ℏ
𝑛
where we have introduced the notation
They represent the “starting” state of the system at time 𝑡 = 0. We note now that, if we know
the starting state, and the perturbing potential and the unperturbed eigenvalues and
(1)
eigenfunctions, we can integrate (𝑎̇ 𝑞 (𝑡)) to obtain the first order, time-dependent correction
(1)
𝑎𝑞 (𝑡), to the expansion coefficients.
then we know the new wavefunction and can calculate the behavior of the system from this
new wavefunction. We can proceed to higher order in this time-dependent perturbation theory.
In general, equating powers of progressively higher order, we obtain
(𝑝+1) 1 (𝑝)
𝑎̇ 𝑞 (𝑡) = ̂𝑝 (𝑡)|ψ𝑛 ⟩
∑ 𝑎𝑛 𝑒𝑥𝑝(𝑖𝜓𝑞𝑛 𝑡) ⟨𝜓𝑞 |𝐻
𝑖ℏ
𝑛
We see that this perturbation theory is also a method of successive approximations, just like
the time-independent perturbation theory. We calculate each higher order correction from the
preceding correction.
Just as for the time-independent perturbation theory, the time-dependent theory is often most
useful for calculating some process to the lowest non-zero order.
Higher order time-dependent perturbation theory is very useful, for example, for understanding
nonlinear optical processes.
First order time-dependent perturbation theory gives the ordinary, linear optical properties of
materials. Higher order time-dependent perturbation theory is used to calculate processes such
as second harmonic generation and two photon absorption in nonlinear optics, processes that
are seen routinely with the high intensities of modern lasers.