Quantum Mechanics Note Date: 2024/05/27
Last Updated: 2024-08-16T15:42:43.994Z
Categories:
Physics Tags:
Physics , Quantum Mechanics Read Time: 12 minutes
A random variable X X X is a function that maps the sample space Ω \Omega Ω to the real number line R \mathbb{R} R .
The probability density function f ( x ) f(x) f ( x ) of a random variable X X X is a function that describes the likelihood of the random variable to take on a specific value.
The cumulative distribution function F ( x ) F(x) F ( x ) of a random variable X X X is a function that describes the probability that the random variable takes on a value less than or equal to x x x .
F ( x ) = P ( X ≤ x ) = ∫ − ∞ x f ( x ) d x . F(x) = P(X \leq x) = \int_{-\infty}^x f(x) dx. F ( x ) = P ( X ≤ x ) = ∫ − ∞ x f ( x ) d x .
The expectation of a random variable X X X is the average value of the random variable.
< X > = ∫ − ∞ ∞ x f ( x ) d x . <X> = \int_{-\infty}^{\infty} x f(x) dx. < X >= ∫ − ∞ ∞ x f ( x ) d x .
The variance of a random variable X X X is a measure of how much the values of the random variable vary.
Δ X 2 = < ( X − < X > ) 2 > = < X 2 > − < X > 2 . \Delta X^2 = <(X - <X>)^2> = <X^2> - <X>^2. Δ X 2 =< ( X − < X > ) 2 >=< X 2 > − < X > 2 .
The standard deviation of a random variable X X X is the square root of the variance.
Δ X = Δ X 2 . \Delta X = \sqrt{\Delta X^2}. Δ X = Δ X 2 .
The probability amplitude ψ ( x ) \psi(x) ψ ( x ) of a random variable X X X is a complex value function that the likelihood of the random variable to take on value x x x is given by ∣ ψ ( x ) ∣ 2 |\psi(x)|^2 ∣ ψ ( x ) ∣ 2 .
In other words, the probability density function f ( x ) f(x) f ( x ) is given by
f ( x ) = ∣ ψ ( x ) ∣ 2 . f(x) = |\psi(x)|^2. f ( x ) = ∣ ψ ( x ) ∣ 2 .
The Gaussian integral is given by
∫ − ∞ ∞ e − x 2 σ 2 d x = π σ . \int_{-\infty}^{\infty} e^{-\frac{x^2}{\sigma^2}} dx = \sqrt{\pi}\sigma. ∫ − ∞ ∞ e − σ 2 x 2 d x = π σ .
The Gaussian integral with odd powers of x x x is given by
∫ − ∞ ∞ x 2 n + 1 e − x 2 σ 2 d x = 0. \int_{-\infty}^{\infty} x^{2n+1} e^{-\frac{x^2}{\sigma^2}} dx = 0. ∫ − ∞ ∞ x 2 n + 1 e − σ 2 x 2 d x = 0.
As this is an odd function, the integral is zero.
The even powers of x x x in Gaussian integral can be calculated using Feynman's trick.
∫ − ∞ ∞ x 2 n e − a x 2 d x = ∫ − ∞ ∞ ( − ∂ ∂ a ) n e − a x 2 = ( − ∂ ∂ a ) n ∫ − ∞ ∞ e − a x 2 d x = ( − ∂ ∂ a ) n π a \begin{align}
\int_{-\infty}^{\infty} x^{2n} e^{-ax^2} dx &=
\int_{-\infty}^{\infty} \left(-\frac{\partial}{\partial a}\right)^n e^{-ax^2} \\
&= \left(-\frac{\partial}{\partial a}\right)^n \int_{-\infty}^{\infty} e^{-ax^2} dx \\
&= \left(-\frac{\partial}{\partial a}\right)^n \sqrt{\frac{\pi}{a}}
\end{align} ∫ − ∞ ∞ x 2 n e − a x 2 d x = ∫ − ∞ ∞ ( − ∂ a ∂ ) n e − a x 2 = ( − ∂ a ∂ ) n ∫ − ∞ ∞ e − a x 2 d x = ( − ∂ a ∂ ) n a π
We can postulate that the state of a particle is described by a complex value wave function Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) , where x x x is the position and t t t is the time.
The probability density to find the particle at position x x x is given by ∣ Ψ ( x , t ) ∣ 2 |\Psi(x, t)|^2 ∣Ψ ( x , t ) ∣ 2 .
And thus, the probability to find the particle in an measurable area A A A is given by
P ( A ) = ∫ A ∣ Ψ ( x , t ) ∣ 2 d x . P(A) = \int_A |\Psi(x, t)|^2 dx. P ( A ) = ∫ A ∣Ψ ( x , t ) ∣ 2 d x .
The wave function Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) must be continuous and differentiable.
The wave function Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) defined on space X X X is normalized if
∫ X ∣ Ψ ( x , t ) ∣ 2 d x = 1. \int_{X} |\Psi(x, t)|^2 dx = 1. ∫ X ∣Ψ ( x , t ) ∣ 2 d x = 1.
If the wave function is not normalized, and if A = ∫ X ∣ Ψ ( x , t ) ∣ 2 d x A=\int_{X} |\Psi(x, t)|^2 dx A = ∫ X ∣Ψ ( x , t ) ∣ 2 d x
is finite and greater than 0 0 0 , then the normalized wave function is given by
Ψ normalized ( x , t ) = Ψ ( x , t ) A . \Psi_{\text{normalized}}(x, t) = \frac{\Psi(x, t)}{\sqrt{A}}. Ψ normalized ( x , t ) = A Ψ ( x , t ) .
If Ψ 1 ( x , t ) \Psi_1(x, t) Ψ 1 ( x , t ) and Ψ 2 ( x , t ) \Psi_2(x, t) Ψ 2 ( x , t ) are two wave functions, then the superposition of the wave functions is given by
Ψ ( x , t ) = c 1 Ψ 1 ( x , t ) + c 2 Ψ 2 ( x , t ) , \Psi(x, t) = c_1 \Psi_1(x, t) + c_2 \Psi_2(x, t), Ψ ( x , t ) = c 1 Ψ 1 ( x , t ) + c 2 Ψ 2 ( x , t ) ,
where c 1 c_1 c 1 and c 2 c_2 c 2 are complex numbers.
And Ψ \Psi Ψ is also a wave function.
The interference of wave functions is a phenomenon where two wave functions Ψ 1 ( x , t ) \Psi_1(x, t) Ψ 1 ( x , t ) and Ψ 2 ( x , t ) \Psi_2(x, t) Ψ 2 ( x , t ) interfere with each other to form a new wave function Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) .
Let Ψ ( x , t ) = Ψ 1 ( x , t ) + Ψ 2 ( x , t ) \Psi(x, t) = \Psi_1(x, t) + \Psi_2(x, t) Ψ ( x , t ) = Ψ 1 ( x , t ) + Ψ 2 ( x , t ) .
Then the probability density of the new wave function is given by
∣ Ψ ( x , t ) ∣ 2 = ∣ Ψ 1 ( x , t ) ∣ 2 + ∣ Ψ 2 ( x , t ) ∣ 2 + Ψ 1 ( x , t ) ∗ Ψ 2 ( x , t ) + Ψ 1 ( x , t ) Ψ 2 ( x , t ) ∗ |\Psi(x, t)|^2 = |\Psi_1(x, t)|^2 + |\Psi_2(x, t)|^2 + \Psi_1(x, t)^* \Psi_2(x, t) + \Psi_1(x, t) \Psi_2(x, t)^* ∣Ψ ( x , t ) ∣ 2 = ∣ Ψ 1 ( x , t ) ∣ 2 + ∣ Ψ 2 ( x , t ) ∣ 2 + Ψ 1 ( x , t ) ∗ Ψ 2 ( x , t ) + Ψ 1 ( x , t ) Ψ 2 ( x , t ) ∗
And the Ψ 1 ( x , t ) ∗ Ψ 2 ( x , t ) + Ψ 1 ( x , t ) Ψ 2 ( x , t ) ∗ \Psi_1(x, t)^* \Psi_2(x, t) + \Psi_1(x, t) \Psi_2(x, t)^* Ψ 1 ( x , t ) ∗ Ψ 2 ( x , t ) + Ψ 1 ( x , t ) Ψ 2 ( x , t ) ∗
is usually called the interference term.
The classical plane wave is given by
Ψ ( x , t ) = A e i k x − i ω t \Psi(x, t) = A e^{ikx - i\omega t} Ψ ( x , t ) = A e ik x − iω t
where A A A is the amplitude, k k k is the wave number, and ω \omega ω is the angular frequency.
Note: The classical plane wave is not normalizable, as the integral of the probability density is infinite. In this case, we can use Ψ ( x , t ) = A e i k x − i ω t − i a x 2 \Psi(x, t) = A e^{ikx - i\omega t -iax^2} Ψ ( x , t ) = A e ik x − iω t − ia x 2 where a a a is a relatively small positive constant. The new wave function is normalizable and have similar behaviour to the plane wave near the origin.
The reduced Planck constant is given by
ℏ = h 2 π ≈ 1.0545718 × 1 0 − 34 J s . \hbar = \frac{h}{2\pi} \approx 1.0545718 \times 10^{-34} \text{ J s}. ℏ = 2 π h ≈ 1.0545718 × 1 0 − 34 J s .
where h h h is the Planck constant.
The quantization rule is given by
p ^ = − i ℏ ∇ E ^ = i ℏ ∂ ∂ t \begin{align}
\hat{p} &= -i \hbar \nabla \\
\hat{E} &= i \hbar \frac{\partial}{\partial t}
\end{align} p ^ E ^ = − i ℏ∇ = i ℏ ∂ t ∂
In classical mechanics, the energy of a free particle is given by
E = T + V = p 2 2 m + V ( x ) . E = T + V = \frac{p^2}{2m} + V(x). E = T + V = 2 m p 2 + V ( x ) .
By applying the quantization rule, the Schrödinger equation (SE) for a free particle with wave equation ψ \psi ψ is given by
E ^ ψ = [ p ^ 2 2 m + V ] ψ i ℏ ∂ ψ ∂ t = [ − ℏ 2 2 m ∇ 2 + V ] ψ = [ − ℏ 2 2 m Δ + V ] ψ \begin{align}
\hat{E}\psi &= \left[\frac{\hat{p}^2}{2m} + V\right]\psi \\
i \hbar \frac{\partial \psi}{\partial t} &= \left[-\frac{\hbar^2}{2m} \nabla^2 + V\right]\psi = \left[-\frac{\hbar^2}{2m} \Delta + V\right]\psi
\end{align} E ^ ψ i ℏ ∂ t ∂ ψ = [ 2 m p ^ 2 + V ] ψ = [ − 2 m ℏ 2 ∇ 2 + V ] ψ = [ − 2 m ℏ 2 Δ + V ] ψ
The Hamiltonian operator H ^ \hat{H} H ^ is given by
H ^ = p ^ 2 2 m + V ( x ) \hat{H} = \frac{\hat{p}^2}{2m} + V(x) H ^ = 2 m p ^ 2 + V ( x )
Using the Hamiltonian operator, the Schrödinger equation for a free particle is given by
E ^ ∂ ψ ∂ t = H ^ ψ . \hat{E} \frac{\partial \psi}{\partial t} = \hat{H}\psi. E ^ ∂ t ∂ ψ = H ^ ψ .
The Schrödinger equation can be solved using separation of variables.
Let ψ ( x , t ) = ϕ ( x ) τ ( t ) \psi(x, t) = \phi(x) \tau(t) ψ ( x , t ) = ϕ ( x ) τ ( t ) .
Then the Schrödinger equation becomes
E ^ ψ = H ^ ψ ϕ E ^ τ = τ H ^ ϕ E ^ τ τ = H ^ ϕ ϕ = E i ℏ ∂ ∂ t τ = E τ τ = e − i E t / ℏ H ^ ϕ = E ϕ \begin{align}
\hat{E}\psi &= \hat{H}\psi \\
\phi \hat{E} \tau &= \tau \hat{H} \phi \\
\frac{\hat{E} \tau}{\tau} &= \frac{\hat{H} \phi}{\phi} = E \\
i\hbar \frac{\partial}{\partial t} \tau &= E \tau \\
\tau &= e^{-iEt/\hbar} \\
\hat{H} \phi &= E \phi
\end{align} E ^ ψ ϕ E ^ τ τ E ^ τ i ℏ ∂ t ∂ τ τ H ^ ϕ = H ^ ψ = τ H ^ ϕ = ϕ H ^ ϕ = E = E τ = e − i Et /ℏ = Eϕ
Assume, ϕ 1 , ϕ 2 , … \phi_1,\phi_2,\ldots ϕ 1 , ϕ 2 , … solve H ^ ϕ = E ϕ \hat{H} \phi = E \phi H ^ ϕ = Eϕ with E 1 , E 2 , … E_1,E_2,\ldots E 1 , E 2 , … .
Then the general solution is given by
ϕ ( x ) = ∑ n = 1 ∞ c n ϕ n ( x ) e − i E n t / ℏ . \phi(x) = \sum_{n=1}^{\infty} c_n \phi_n(x)e^{-iE_nt/\hbar}. ϕ ( x ) = n = 1 ∑ ∞ c n ϕ n ( x ) e − i E n t /ℏ .
By the previous discussion , the task of solving the Schrödinger equation is reduced to solving the equation
H ^ ϕ n = E n ϕ n . \hat{H} \phi_n = E_n \phi_n. H ^ ϕ n = E n ϕ n .
This equation is called the time-independent Schrödinger equation (TISE) .
We always call E n E_n E n eigenvalues of H H H . And it is also known as:
Energy lLevels
Eigenenergies
Energy Eigenvalues
And ϕ n \phi_n ϕ n are called energy eigenstates of H H H . And it is also known as:
Stationary States
Energy Eigenfunctions
If two or more energy eigenstates have the same energy eigenvalue, then the energy eigenvalue is said to be degenerate.
If two eigenstates have the same eigenvalue, we call it degeneracy of order 2 or twice degenerate .
Consider the following double slit experiment setup.
We can assume the wave function that go through the slits A A A and B B B are the same,
and is given by the analogue of plane wave ψ ( r , t ) = e i k x − i w t r \psi(r, t) = \frac{e^{ikx-iwt}}{r} ψ ( r , t ) = r e ik x − i wt .
Then, for any point P P P on the screen, the wave function is the superposition of the wave functions from the slits A A A and B B B .
Ψ ( x , t ) = ψ A ( r 1 , t ) + ψ B ( r 2 , t ) = e i k r 1 − i w t + e i k r 2 − i w t = e − i w t [ e i k r 1 r 1 + e i k r 2 r 2 ] \begin{align}
\Psi(x, t) &= \psi_A(r_1, t) + \psi_B(r_2, t) \\
&= e^{ikr_1-iwt} + e^{ikr_2-iwt} \\
&= e^{-iwt}\left[\frac{e^{ikr_1}}{r_1} + \frac{e^{ikr_2}}{r_2}\right]
\end{align} Ψ ( x , t ) = ψ A ( r 1 , t ) + ψ B ( r 2 , t ) = e ik r 1 − i wt + e ik r 2 − i wt = e − i wt [ r 1 e ik r 1 + r 2 e ik r 2 ]
where r 1 r_1 r 1 and r 2 r_2 r 2 are the distances from the slits A A A and B B B to the point P P P .
And is given by
r 1 = L 2 + ( x − d ) 2 r 2 = L 2 + ( x + d ) 2 \begin{align}
r_1 &= \sqrt{L^2 + (x-d)^2} \\
r_2 &= \sqrt{L^2 + (x+d)^2}
\end{align} r 1 r 2 = L 2 + ( x − d ) 2 = L 2 + ( x + d ) 2
Without normalization, the probability density of the wave function is proportional to
∣ Ψ ( x , t ) ∣ 2 = ∣ e i k r 1 r 1 + e i k r 2 r 2 ∣ 2 = 1 r 1 2 + 1 r 2 2 + 1 r 1 r 2 [ e i k ( r 1 − r 2 ) + e − i k ( r 1 − r 2 ) ] ≈ 1 L 2 [ 2 + 2 cos ( k ( r 1 − r 2 ) ) ] = 1 L 2 [ 2 + 2 cos ( k ( L 2 + ( x + d ) 2 − L 2 + ( x − d ) 2 ) ) ] = 1 L 2 [ 2 + 2 cos ( k 4 d x L 2 + ( x + d ) 2 + L 2 + ( x − d ) 2 ) ] ≈ 1 L 2 [ 2 + 2 cos ( k 4 d x 2 L ) ] As L ≫ x , d = 1 L 2 [ 2 + 2 cos ( k 2 d x L ) ] \begin{align}
|\Psi(x, t)|^2 &= \left|\frac{e^{ikr_1}}{r_1} + \frac{e^{ikr_2}}{r_2}\right|^2 \\
&= \frac{1}{r_1^2} + \frac{1}{r_2^2} + \frac{1}{r_1r_2}\left[e^{ik(r_1-r_2)} + e^{-ik(r_1-r_2)}\right] \\
&\approx \frac{1}{L^2}\left[2 + 2\cos(k(r_1-r_2))\right] \\
&= \frac{1}{L^2}\left[2 + 2\cos\left(
k
\left(
\sqrt{L^2 + (x+d)^2} - \sqrt{L^2 + (x-d)^2}
\right)
\right) \right] \\
&= \frac{1}{L^2}\left[2 + 2\cos\left(
k
\frac{
4dx
}{
\sqrt{L^2 + (x+d)^2} + \sqrt{L^2 + (x-d)^2}
}
\right) \right] \\
&\approx \frac{1}{L^2}\left[2 + 2\cos\left(
k
\frac{
4dx
}{
2L
}
\right)\right] \quad\text{As } L \gg x,d \\
&= \frac{1}{L^2}\left[2 + 2\cos\left(
k
\frac{
2dx
}{
L
}
\right)\right]
\end{align} ∣Ψ ( x , t ) ∣ 2 = r 1 e ik r 1 + r 2 e ik r 2 2 = r 1 2 1 + r 2 2 1 + r 1 r 2 1 [ e ik ( r 1 − r 2 ) + e − ik ( r 1 − r 2 ) ] ≈ L 2 1 [ 2 + 2 cos ( k ( r 1 − r 2 )) ] = L 2 1 [ 2 + 2 cos ( k ( L 2 + ( x + d ) 2 − L 2 + ( x − d ) 2 ) ) ] = L 2 1 [ 2 + 2 cos ( k L 2 + ( x + d ) 2 + L 2 + ( x − d ) 2 4 d x ) ] ≈ L 2 1 [ 2 + 2 cos ( k 2 L 4 d x ) ] As L ≫ x , d = L 2 1 [ 2 + 2 cos ( k L 2 d x ) ]
We can see that the probability density of the wave function is proportional to a transformed cosine function, which explains the interference pattern on the screen.
Given a potential V ( x ) : R → C V(x): \mathbb{R}\rightarrow \mathbb{C} V ( x ) : R → C , the potential is said to be even if
V ( x ) = V ( − x ) V(x) = V(-x) V ( x ) = V ( − x )
In such cases, we can assume the TISE solution has definite parity , that is:
ϕ ( x ) = ± ϕ ( − x ) \phi(x) = \pm \phi(-x) ϕ ( x ) = ± ϕ ( − x )
Given a real potential V ( x ) V(x) V ( x ) , we can assume the TISE solution to have definite parity regarding conjugation , that is:
ϕ ( x ) = ϕ ∗ ( x ) \phi(x) = \phi^*(x) ϕ ( x ) = ϕ ∗ ( x )
In general there are three kinds of states of TISE solutions:
Bound States
Scattering States
Non-physical States
Bound states are states where the energy eigenvalue E n E_n E n is less than the potential energy at infinity and greater than the minimum potential energy.
In such cases, the TISE solution is normalizable, and the energy eigenvalue is quantized.
Scattering states are states where the energy eigenvalue E n E_n E n is greater than the potential energy at infinity.
In such cases, the TISE solution is not normalizable, and the energy eigenvalue is continuous.
Scattering states can be approximated by normalizable states by similar method as the classical plane wave .
Non-physical states are states where the energy eigenvalue E n E_n E n is less than the minimum potential energy.
This case cannot happen in real physical systems.
Consider the potential to be real. Define the probability density of a wave function P = Ψ ( x , t ) P = \Psi(x, t) P = Ψ ( x , t ) as ∣ Ψ ( x , t ) ∣ 2 |\Psi(x, t)|^2 ∣Ψ ( x , t ) ∣ 2 .
Then:
∂ ∂ t P = ∂ ψ ∂ t ψ ∗ + ψ ∂ ψ ∂ t ∗ = 1 i ℏ [ H ^ ψ ψ ∗ − ψ ( H ^ ψ ) ∗ ] = − ℏ 2 i ℏ 2 m [ ∂ 2 ψ ∂ x 2 ψ ∗ − ψ ∂ 2 ψ ∂ x 2 ∗ ] = − d d x [ ℏ 2 i ℏ 2 m [ ∂ ψ ∂ x ψ ∗ − ∂ ψ ∂ x ∂ ψ ∂ x ∗ − ψ ∂ ψ ∂ x ∗ + ∂ ψ ∂ x ∂ ψ ∂ x ∗ ] ] = − d d x [ ℏ 2 m i [ − ψ ∂ ψ ∂ x ∗ + ψ ∗ ∂ ψ ∂ x ] ] \begin{align}
\frac{\partial}{\partial t} P &=
\frac{\partial \psi}{\partial t} \psi^* + \psi \frac{\partial \psi}{\partial t}^* \\
&= \frac{1}{i\hbar}\left[
\hat{H}\psi\psi^* - \psi(\hat{H}\psi)^*
\right] \\
&= -\frac{\hbar^2}{i\hbar 2m}\left[
\frac{\partial^2 \psi}{\partial x^2}\psi^* - \psi\frac{\partial^2 \psi}{\partial x^2}^*
\right] \\
&= -\frac{d}{dx}\left[\frac{\hbar^2}{i\hbar 2m}\left[
\frac{\partial \psi}{\partial x}\psi^*
- \frac{\partial \psi}{\partial x}\frac{\partial \psi}{\partial x}^*
- \psi\frac{\partial \psi}{\partial x}^*
+ \frac{\partial \psi}{\partial x}\frac{\partial \psi}{\partial x}^*
\right]\right] \\
&= -\frac{d}{dx}\left[\frac{\hbar}{2mi}\left[
-\psi\frac{\partial \psi}{\partial x}^* + \psi^*\frac{\partial \psi}{\partial x}
\right]\right]
\end{align} ∂ t ∂ P = ∂ t ∂ ψ ψ ∗ + ψ ∂ t ∂ ψ ∗ = i ℏ 1 [ H ^ ψ ψ ∗ − ψ ( H ^ ψ ) ∗ ] = − i ℏ2 m ℏ 2 [ ∂ x 2 ∂ 2 ψ ψ ∗ − ψ ∂ x 2 ∂ 2 ψ ∗ ] = − d x d [ i ℏ2 m ℏ 2 [ ∂ x ∂ ψ ψ ∗ − ∂ x ∂ ψ ∂ x ∂ ψ ∗ − ψ ∂ x ∂ ψ ∗ + ∂ x ∂ ψ ∂ x ∂ ψ ∗ ] ] = − d x d [ 2 mi ℏ [ − ψ ∂ x ∂ ψ ∗ + ψ ∗ ∂ x ∂ ψ ] ]
We thus define the probability current J J J as
J = ℏ 2 m i [ − ψ ∂ ψ ∂ x ∗ + ψ ∗ ∂ ψ ∂ x ] J = \frac{\hbar}{2mi}\left[
-\psi\frac{\partial \psi}{\partial x}^* + \psi^*\frac{\partial \psi}{\partial x}
\right] J = 2 mi ℏ [ − ψ ∂ x ∂ ψ ∗ + ψ ∗ ∂ x ∂ ψ ]
The above equation become:
∂ P ∂ t = − d J d x \frac{\partial P}{\partial t} = -\frac{dJ}{dx} ∂ t ∂ P = − d x dJ
and is called the continuity equation .
We consider a potential barrier of height V 0 V_0 V 0 and width 2 a 2a 2 a .
That is, the potential is given by
V ( x ) = { V 0 if ∣ x ∣ < 2 a 0 otherwise V(x) = \begin{cases}
V_0 & \text{if } |x| < 2a \\
0 & \text{otherwise}
\end{cases} V ( x ) = { V 0 0 if ∣ x ∣ < 2 a otherwise
Then, the TISE can be divided into three regions:
Region I: x < − a x < -a x < − a
d 2 ψ d x 2 = − 2 m ℏ 2 E ψ \frac{d^2 \psi}{dx^2} = -\frac{2m}{\hbar^2}E\psi d x 2 d 2 ψ = − ℏ 2 2 m E ψ
Region II: − a < x < a -a < x < a − a < x < a
d 2 ψ d x 2 = − 2 m ℏ 2 ( E − V 0 ) ψ \frac{d^2 \psi}{dx^2} = -\frac{2m}{\hbar^2}(E-V_0)\psi d x 2 d 2 ψ = − ℏ 2 2 m ( E − V 0 ) ψ
Region III: x > a x > a x > a
d 2 ψ d x 2 = − 2 m ℏ 2 E ψ \frac{d^2 \psi}{dx^2} = -\frac{2m}{\hbar^2}E\psi d x 2 d 2 ψ = − ℏ 2 2 m E ψ
For the sake of simplicity,
let k = 2 m E ℏ 2 k = \sqrt{\frac{2mE}{\hbar^2}} k = ℏ 2 2 m E
we can assume the solution of the TISE in Region I and Region III to be
In region I:
ψ ( x ) = A e i k x + B e − i k x \psi(x) = A e^{ikx} + B e^{-ikx} ψ ( x ) = A e ik x + B e − ik x
In region III:
ψ ( x ) = F e i k x \psi(x) = F e^{ikx} ψ ( x ) = F e ik x
The wave A e i k x A e^{ikx} A e ik x can be interpreted as the wave go toward the barrier, and the wave B e − i k x B e^{-ikx} B e − ik x can be interpreted as the wave being reflected by the barrier and wave F e i k x F e^{ikx} F e ik x can be interpreted as the wave go through the barrier.
As there is no probability for the particle to be in region II, the probability current must be same at the point x = a x = a x = a and x = − a x = -a x = − a .
In region I:
J = ℏ 2 m i [ ψ − ∂ ψ ∂ x ∗ + ψ ∗ ∂ ψ ∂ x ] = ℏ k m ( ∣ A ∣ 2 − ∣ B ∣ 2 ) J = \frac{\hbar}{2mi}\left[
\psi\frac{-\partial \psi}{\partial x}^* + \psi^*\frac{\partial \psi}{\partial x}
\right] = \frac{\hbar k}{m}(|A|^2 - |B|^2) J = 2 mi ℏ [ ψ ∂ x − ∂ ψ ∗ + ψ ∗ ∂ x ∂ ψ ] = m ℏ k ( ∣ A ∣ 2 − ∣ B ∣ 2 )
In region III:
J = ℏ 2 m i [ ψ − ∂ ψ ∂ x ∗ + ψ ∗ ∂ ψ ∂ x ] = ℏ k m ∣ F ∣ 2 J = \frac{\hbar}{2mi}\left[
\psi\frac{-\partial \psi}{\partial x}^* + \psi^*\frac{\partial \psi}{\partial x}
\right] = \frac{\hbar k}{m}|F|^2 J = 2 mi ℏ [ ψ ∂ x − ∂ ψ ∗ + ψ ∗ ∂ x ∂ ψ ] = m ℏ k ∣ F ∣ 2
Thus, we have
∣ A ∣ 2 − ∣ B ∣ 2 = ∣ F ∣ 2 |A|^2 - |B|^2 = |F|^2 ∣ A ∣ 2 − ∣ B ∣ 2 = ∣ F ∣ 2
We define the reflection probability R R R and transmission probability T T T as
R = ∣ B ∣ 2 ∣ A ∣ 2 T = ∣ F ∣ 2 ∣ A ∣ 2 \begin{align}
R &= \frac{|B|^2}{|A|^2} \\
T &= \frac{|F|^2}{|A|^2}
\end{align} R T = ∣ A ∣ 2 ∣ B ∣ 2 = ∣ A ∣ 2 ∣ F ∣ 2
Then, we have
R + T = 1 R + T = 1 R + T = 1
We consider a barrier that is infinitely high and thin at origin.
To state this formally, we consider the potential to be the limiting behaviour of
the Dirac delta function potential.
δ L ( x ) = { 1 L if − L 2 < x < L 2 0 otherwise \delta_L(x) = \begin{cases}
\frac{1}{L} & \text{if } -\frac{L}{2} < x < \frac{L}{2} \\
0 & \text{otherwise}
\end{cases} δ L ( x ) = { L 1 0 if − 2 L < x < 2 L otherwise
and
δ = lim L → 0 δ L ( x ) \delta = \lim_{L \to 0} \delta_L(x) δ = L → 0 lim δ L ( x )
Also, for all function g ( x ) g(x) g ( x ) , we have
lim ϵ → 0 ∫ − ϵ ϵ g ( x ) δ ( x ) d x = g ( 0 ) \lim_{\epsilon \to 0} \int_{-\epsilon}^{\epsilon} g(x) \delta(x) dx = g(0) ϵ → 0 lim ∫ − ϵ ϵ g ( x ) δ ( x ) d x = g ( 0 )
Then, the potential is given by
V ( x ) = α δ ( x ) V(x) = \alpha \delta(x) V ( x ) = α δ ( x )
We again, consider the simplified model in the previous section .
In region I:
ψ ( x ) = A e i k x + B e − i k x \psi(x) = A e^{ikx} + B e^{-ikx} ψ ( x ) = A e ik x + B e − ik x
In region III:
ψ ( x ) = F e i k x \psi(x) = F e^{ikx} ψ ( x ) = F e ik x
As the region II is infinity thin,
the right hand side of the region I and the left hand side of the region III must be the same.
Thus, we have
A + B = F A + B = F A + B = F
Also, by integrating the TISE over the region [ − ϵ , ϵ ] [-\epsilon,\epsilon] [ − ϵ , ϵ ] ,
we get:
∫ − ϵ ϵ d 2 ψ d x 2 d x = − 2 m ℏ 2 ∫ − ϵ ϵ ( E − α δ ( x ) ) ψ d x [ d ψ d x ] − ϵ ϵ = − 2 m ℏ 2 [ E ψ ( ϵ ) − E ψ ( − ϵ ) − α ∫ − ϵ ϵ ψ δ d x ] \begin{align}
\int_{-\epsilon}^{\epsilon} \frac{d^2 \psi}{dx^2} dx &= -\frac{2m}{\hbar^2} \int_{-\epsilon}^{\epsilon} (E-\alpha\delta(x))\psi dx \\
\left[\frac{d \psi}{dx}\right]_{-\epsilon}^{\epsilon} &= -\frac{2m}{\hbar^2}
\left[E\psi(\epsilon)-E\psi(-\epsilon) - \alpha\int_{-\epsilon}^{\epsilon} \psi\delta dx\right] \\
\end{align} ∫ − ϵ ϵ d x 2 d 2 ψ d x [ d x d ψ ] − ϵ ϵ = − ℏ 2 2 m ∫ − ϵ ϵ ( E − α δ ( x )) ψ d x = − ℏ 2 2 m [ E ψ ( ϵ ) − E ψ ( − ϵ ) − α ∫ − ϵ ϵ ψ δ d x ]
Taking the limit ϵ → 0 \epsilon \to 0 ϵ → 0 , we get
[ d ψ d x ] 0 + − [ d ψ d x ] 0 − = 2 m ℏ 2 α ψ ( 0 ) \begin{align}
\left[\frac{d \psi}{dx}\right]_{0^+} - \left[\frac{d \psi}{dx}\right]_{0^-} &= \frac{2m}{\hbar^2} \alpha \psi(0)
\end{align} [ d x d ψ ] 0 + − [ d x d ψ ] 0 − = ℏ 2 2 m α ψ ( 0 )
Substituting the solution of the TISE in region I and region III, we get
i k ( A − B ) − i k F = 2 m ℏ 2 α F ik(A-B) - ikF = \frac{2m}{\hbar^2} \alpha F ik ( A − B ) − ik F = ℏ 2 2 m α F
Solving the above equation, we can get the reflection and transmission probability.
Define the Hilbert space H \mathbf{H} H as the space of all possible wave functions Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) that are square integrable.
H = { Ψ ( x , t ) ∣ ∫ − ∞ ∞ ∣ Ψ ( x , t ) ∣ 2 d x < ∞ } \mathbf{H} = \left\{
\Psi(x, t) \mid \int_{-\infty}^{\infty} |\Psi(x, t)|^2 dx < \infty
\right\} H = { Ψ ( x , t ) ∣ ∫ − ∞ ∞ ∣Ψ ( x , t ) ∣ 2 d x < ∞ }
The Hilbert space is a complex vector space .
We define the inner product of two wave functions Ψ 1 ( x , t ) \Psi_1(x, t) Ψ 1 ( x , t ) and Ψ 2 ( x , t ) \Psi_2(x, t) Ψ 2 ( x , t ) as
< Ψ 1 , Ψ 2 > = ∫ − ∞ ∞ Ψ 1 ∗ ( x , t ) Ψ 2 ( x , t ) d x \left<\Psi_1, \Psi_2\right> = \int_{-\infty}^{\infty} \Psi_1^*(x, t) \Psi_2(x, t) dx ⟨ Ψ 1 , Ψ 2 ⟩ = ∫ − ∞ ∞ Ψ 1 ∗ ( x , t ) Ψ 2 ( x , t ) d x
Linearity in Second Argument : For all Ψ 1 , Ψ 2 ∈ H \Psi_1, \Psi_2 \in \mathbf{H} Ψ 1 , Ψ 2 ∈ H and a , b ∈ C a,b \in \mathbb{C} a , b ∈ C , we have < a Ψ 1 + b Ψ 2 , Ψ 3 > = a < Ψ 1 , Ψ 3 > + b < Ψ 2 , Ψ 3 > \left<a\Psi_1 + b\Psi_2, \Psi_3\right> = a\left<\Psi_1, \Psi_3\right> + b\left<\Psi_2, \Psi_3\right> ⟨ a Ψ 1 + b Ψ 2 , Ψ 3 ⟩ = a ⟨ Ψ 1 , Ψ 3 ⟩ + b ⟨ Ψ 2 , Ψ 3 ⟩ .
Anti-linearity in First Argument : For all Ψ 1 , Ψ 2 ∈ H \Psi_1, \Psi_2 \in \mathbf{H} Ψ 1 , Ψ 2 ∈ H and a , b ∈ C a,b \in \mathbb{C} a , b ∈ C , we have < Ψ 1 , a Ψ 2 + b Ψ 3 > = a ∗ < Ψ 1 , Ψ 2 > + b ∗ < Ψ 1 , Ψ 3 > \left<\Psi_1, a\Psi_2 + b\Psi_3\right> = a^*\left<\Psi_1, \Psi_2\right> + b^*\left<\Psi_1, \Psi_3\right> ⟨ Ψ 1 , a Ψ 2 + b Ψ 3 ⟩ = a ∗ ⟨ Ψ 1 , Ψ 2 ⟩ + b ∗ ⟨ Ψ 1 , Ψ 3 ⟩ .
Positive Definite : For all Ψ ∈ H \Psi \in \mathbf{H} Ψ ∈ H , we have < Ψ , Ψ > ≥ 0 \left<\Psi, \Psi\right> \geq 0 ⟨ Ψ , Ψ ⟩ ≥ 0 and < Ψ , Ψ > = 0 \left<\Psi, \Psi\right> = 0 ⟨ Ψ , Ψ ⟩ = 0 if and only if Ψ = 0 \Psi = 0 Ψ = 0 .
Conjugate Symmetric (Skew Symmetric) : For all Ψ 1 , Ψ 2 ∈ H \Psi_1, \Psi_2 \in \mathbf{H} Ψ 1 , Ψ 2 ∈ H , we have < Ψ 1 , Ψ 2 > = < Ψ 2 , Ψ 1 > ∗ \left<\Psi_1, \Psi_2\right> = \left<\Psi_2, \Psi_1\right>^* ⟨ Ψ 1 , Ψ 2 ⟩ = ⟨ Ψ 2 , Ψ 1 ⟩ ∗ .
The norm of a wave function Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) is defined as
∣ ∣ Ψ ∣ ∣ = < Ψ , Ψ > ||\Psi|| = \sqrt{\left<\Psi, \Psi\right>} ∣∣Ψ∣∣ = ⟨ Ψ , Ψ ⟩
Two wave functions Ψ 1 ( x , t ) \Psi_1(x, t) Ψ 1 ( x , t ) and Ψ 2 ( x , t ) \Psi_2(x, t) Ψ 2 ( x , t ) are said to be orthogonal if
< Ψ 1 , Ψ 2 > = 0 \left<\Psi_1, \Psi_2\right> = 0 ⟨ Ψ 1 , Ψ 2 ⟩ = 0
The angle between two wave functions Ψ 1 ( x , t ) \Psi_1(x, t) Ψ 1 ( x , t ) and Ψ 2 ( x , t ) \Psi_2(x, t) Ψ 2 ( x , t ) is defined as
cos θ = < Ψ 1 , Ψ 2 > ∣ ∣ Ψ 1 ∣ ∣ ∣ ∣ Ψ 2 ∣ ∣ \cos\theta = \frac{\left<\Psi_1, \Psi_2\right>}{||\Psi_1|| ||\Psi_2||} cos θ = ∣∣ Ψ 1 ∣∣∣∣ Ψ 2 ∣∣ ⟨ Ψ 1 , Ψ 2 ⟩
A set of wave functions { Ψ 1 ( x , t ) , Ψ 2 ( x , t ) , … } \{\Psi_1(x, t), \Psi_2(x, t), \ldots\} { Ψ 1 ( x , t ) , Ψ 2 ( x , t ) , … } is said to be an orthonormal basis if
The set is orthogonal.
The set is normalized.
The set spans the Hilbert space. That is, for all Ψ ( x , t ) ∈ H \Psi(x, t) \in \mathbf{H} Ψ ( x , t ) ∈ H , there exists a set of complex numbers { c 1 , c 2 , … } \{c_1, c_2, \ldots\} { c 1 , c 2 , … } such that
Ψ ( x , t ) = ∑ n = 1 ∞ c n Ψ n ( x , t ) \Psi(x, t) = \sum_{n=1}^{\infty} c_n \Psi_n(x, t) Ψ ( x , t ) = n = 1 ∑ ∞ c n Ψ n ( x , t )
An operator is a function that maps a wave function to another wave function.
The Hermitian conjugate of an operator A ^ \hat{A} A ^ is denoted by A ^ † \hat{A}^\dagger A ^ † and is the unique operator that satisfies
< A ^ Ψ 1 , Ψ 2 > = < Ψ 1 , A ^ † Ψ 2 > \left<\hat{A}\Psi_1, \Psi_2\right> = \left<\Psi_1, \hat{A}^\dagger\Psi_2\right> ⟨ A ^ Ψ 1 , Ψ 2 ⟩ = ⟨ Ψ 1 , A ^ † Ψ 2 ⟩
An Hermitian operator is an operator that satisfies
< A ^ Ψ 1 , Ψ 2 > = < Ψ 1 , A ^ Ψ 2 > \left<\hat{A}\Psi_1, \Psi_2\right> = \left<\Psi_1, \hat{A}\Psi_2\right> ⟨ A ^ Ψ 1 , Ψ 2 ⟩ = ⟨ Ψ 1 , A ^ Ψ 2 ⟩
for all Ψ 1 , Ψ 2 ∈ H \Psi_1, \Psi_2 \in \mathbf{H} Ψ 1 , Ψ 2 ∈ H .
An equivalent definition is that the operator is equal to its Hermitian conjugate.
The spectral theorem states that for all Hermitian operators A ^ \hat{A} A ^ , there exists an orthonormal basis { Ψ 1 ( x , t ) , Ψ 2 ( x , t ) , … } \{\Psi_1(x, t), \Psi_2(x, t), \ldots\} { Ψ 1 ( x , t ) , Ψ 2 ( x , t ) , … } such that
A ^ Ψ n ( x , t ) = a n Ψ n ( x , t ) \hat{A}\Psi_n(x, t) = a_n\Psi_n(x, t) A ^ Ψ n ( x , t ) = a n Ψ n ( x , t )
where a n a_n a n are the eigenvalues of A ^ \hat{A} A ^ and is real.
The Hamiltonian operator H ^ \hat{H} H ^ is Hermitian.
Proof:
< H ^ Ψ 1 , Ψ 2 > = ∫ − ∞ ∞ Ψ 1 ∗ ( x , t ) H ^ Ψ 2 ( x , t ) d x = ∫ − ∞ ∞ Ψ 1 ∗ ( x , t ) [ − ℏ 2 2 m ∂ 2 Ψ 2 ∂ x 2 + V ( x ) Ψ 2 ( x , t ) ] d x By Integration By Part = ∫ − ∞ ∞ [ − ℏ 2 2 m ∂ 2 Ψ 1 ∗ ∂ x 2 + V ( x ) Ψ 1 ∗ ( x , t ) ] Ψ 2 ( x , t ) d x = < Ψ 1 , H ^ Ψ 2 > \begin{align}
\left<\hat{H}\Psi_1, \Psi_2\right> &= \int_{-\infty}^{\infty} \Psi_1^*(x, t) \hat{H}\Psi_2(x, t) dx \\
&= \int_{-\infty}^{\infty} \Psi_1^*(x, t) \left[-\frac{\hbar^2}{2m}\frac{\partial^2 \Psi_2}{\partial x^2} + V(x)\Psi_2(x, t)\right] dx \quad \text{By Integration By Part} \\
&= \int_{-\infty}^{\infty} \left[-\frac{\hbar^2}{2m}\frac{\partial^2 \Psi_1^*}{\partial x^2} + V(x)\Psi_1^*(x, t)\right] \Psi_2(x, t) dx \\
&= \left<\Psi_1, \hat{H}\Psi_2\right>
\end{align} ⟨ H ^ Ψ 1 , Ψ 2 ⟩ = ∫ − ∞ ∞ Ψ 1 ∗ ( x , t ) H ^ Ψ 2 ( x , t ) d x = ∫ − ∞ ∞ Ψ 1 ∗ ( x , t ) [ − 2 m ℏ 2 ∂ x 2 ∂ 2 Ψ 2 + V ( x ) Ψ 2 ( x , t ) ] d x By Integration By Part = ∫ − ∞ ∞ [ − 2 m ℏ 2 ∂ x 2 ∂ 2 Ψ 1 ∗ + V ( x ) Ψ 1 ∗ ( x , t ) ] Ψ 2 ( x , t ) d x = ⟨ Ψ 1 , H ^ Ψ 2 ⟩
An operator A ^ \hat{A} A ^ is said to be positive definite if
give any non-zero wave function Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) , we have
< A ^ Ψ , Ψ > > 0 \left<\hat{A}\Psi, \Psi\right> > 0 ⟨ A ^ Ψ , Ψ ⟩ > 0
An operator A ^ \hat{A} A ^ is said to be positive semi-definite if
give any wave function Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) , we have
< A ^ Ψ , Ψ > ≥ 0 \left<\hat{A}\Psi, \Psi\right> \geq 0 ⟨ A ^ Ψ , Ψ ⟩ ≥ 0
Any operator given by A ^ = B ^ † B ^ \hat{A} = \hat{B}^\dagger \hat{B} A ^ = B ^ † B ^ is positive semi-definite.
Given a orthonormal basis { Ψ 1 ( x , t ) , Ψ 2 ( x , t ) , … } \{\Psi_1(x, t), \Psi_2(x, t), \ldots\} { Ψ 1 ( x , t ) , Ψ 2 ( x , t ) , … } , the measurement postulate states that the probability of measuring the normalized wave function Ψ ( x , t ) \Psi(x, t) Ψ ( x , t ) to be in the state Ψ n ( x , t ) \Psi_n(x, t) Ψ n ( x , t ) is given by
P n = ∣ < Ψ n , Ψ > ∣ 2 P_n = \left|\left<\Psi_n, \Psi\right>\right|^2 P n = ∣ ⟨ Ψ n , Ψ ⟩ ∣ 2
After the measurement, the state of the wave function will collapse to the state that is measured.
Given an observable A ^ \hat{A} A ^ ,
it is postulated that A ^ \hat{A} A ^ is Hermitian and has an orthonormal basis { Ψ 1 ( x , t ) , Ψ 2 ( x , t ) , … } \{\Psi_1(x, t), \Psi_2(x, t), \ldots\} { Ψ 1 ( x , t ) , Ψ 2 ( x , t ) , … } .
The expectation value of the observable A ^ \hat{A} A ^ of a wave function ψ \psi ψ is given by
< A ^ > = ∑ n = 1 ∞ a n P n = < ψ ∣ A ^ ψ > \left<\hat{A}\right> = \sum_{n=1}^{\infty} a_n P_n = \left<\psi|\hat{A}\psi\right> ⟨ A ^ ⟩ = n = 1 ∑ ∞ a n P n = ⟨ ψ ∣ A ^ ψ ⟩
In general, two operators A ^ \hat{A} A ^ and B ^ \hat{B} B ^ do not commute.
We define the commutator (Lie Bracket) of two operators A ^ \hat{A} A ^ and B ^ \hat{B} B ^ as
[ A ^ , B ^ ] = A ^ B ^ − B ^ A ^ [\hat{A}, \hat{B}] = \hat{A}\hat{B} - \hat{B}\hat{A} [ A ^ , B ^ ] = A ^ B ^ − B ^ A ^
Linearity : [ a A ^ + b B ^ , C ^ ] = a [ A ^ , C ^ ] + b [ B ^ , C ^ ] [a\hat{A} + b\hat{B}, \hat{C}] = a[\hat{A}, \hat{C}] + b[\hat{B}, \hat{C}] [ a A ^ + b B ^ , C ^ ] = a [ A ^ , C ^ ] + b [ B ^ , C ^ ]
Anti-linearity : [ A ^ , a B ^ + b C ^ ] = a [ A ^ , B ^ ] + b [ A ^ , C ^ ] [\hat{A}, a\hat{B} + b\hat{C}] = a[\hat{A}, \hat{B}] + b[\hat{A}, \hat{C}] [ A ^ , a B ^ + b C ^ ] = a [ A ^ , B ^ ] + b [ A ^ , C ^ ]
Linearity in Second Argument : [ C ^ , a A ^ + b B ^ ] = a [ C ^ , A ^ ] + b [ C ^ , B ^ ] [\hat{C}, a\hat{A} + b\hat{B}] = a[\hat{C},\hat{A}] + b[\hat{C},\hat{B}] [ C ^ , a A ^ + b B ^ ] = a [ C ^ , A ^ ] + b [ C ^ , B ^ ]
Distributivity : [ A ^ , B ^ C ^ ] = [ A ^ , B ^ ] C ^ + B ^ [ A ^ , C ^ ] [\hat{A}, \hat{B}\hat{C}] = [\hat{A}, \hat{B}]\hat{C} + \hat{B}[\hat{A}, \hat{C}] [ A ^ , B ^ C ^ ] = [ A ^ , B ^ ] C ^ + B ^ [ A ^ , C ^ ]
The commutator of position operator x ^ \hat{x} x ^ and momentum operator p ^ \hat{p} p ^ is given by
[ x ^ , p ^ ] = x ^ p ^ − p ^ x ^ = x ( − i ℏ ∂ ∂ x ) − ( − i ℏ ∂ ∂ x ) x = x ( − i ℏ ∂ ∂ x ) − ( − i ℏ ) − x ( − i ℏ ∂ ∂ x ) = i ℏ \begin{align}
[\hat{x},\hat{p}] &= \hat{x}\hat{p} - \hat{p}\hat{x} \\
&= x\left(-i\hbar\frac{\partial}{\partial x}\right) - \left(-i\hbar\frac{\partial}{\partial x}\right)x \\
&= x\left(-i\hbar\frac{\partial}{\partial x}\right) - \left(-i\hbar\right) - x\left(-i\hbar\frac{\partial}{\partial x}\right) \\
&= i\hbar
\end{align} [ x ^ , p ^ ] = x ^ p ^ − p ^ x ^ = x ( − i ℏ ∂ x ∂ ) − ( − i ℏ ∂ x ∂ ) x = x ( − i ℏ ∂ x ∂ ) − ( − i ℏ ) − x ( − i ℏ ∂ x ∂ ) = i ℏ
Two observables A ^ \hat{A} A ^ and B ^ \hat{B} B ^ are said to be compatible if they commute.
The Robertson Inequality states that for any two observables A ^ \hat{A} A ^ and B ^ \hat{B} B ^ , we have
Δ A ^ Δ B ^ ≥ 1 2 ∣ < [ A ^ , B ^ ] > ∣ \Delta \hat{A} \Delta \hat{B} \ge \frac{1}{2} \left|\left<[\hat{A}, \hat{B}]\right>\right| Δ A ^ Δ B ^ ≥ 2 1 ⟨ [ A ^ , B ^ ] ⟩
Proof:
Without loss of generality, we can assume < A ^ > = 0 \left<\hat{A}\right> = 0 ⟨ A ^ ⟩ = 0 and < B ^ > = 0 \left<\hat{B}\right> = 0 ⟨ B ^ ⟩ = 0 . Otherwise we replace A ^ \hat{A} A ^ with A ^ − < A ^ > \hat{A} - \left<\hat{A}\right> A ^ − ⟨ A ^ ⟩ and B ^ \hat{B} B ^ with B ^ − < B ^ > \hat{B} - \left<\hat{B}\right> B ^ − ⟨ B ^ ⟩ .
Then, we have
Δ A ^ Δ B ^ = < A ^ 2 > < B ^ 2 > = < A ^ ψ ∣ A ^ ψ > < B ^ ψ ∣ B ^ ψ > ≥ ∣ < A ^ ψ ∣ B ^ ψ > ∣ Cauchy-Schwartz = ∣ < ψ ∣ A ^ B ^ ψ > ∣ \begin{align}
\Delta \hat{A} \Delta \hat{B} &= \sqrt{\left<\hat{A}^2\right>\left<\hat{B}^2\right>} \\
&= \sqrt{\left<\hat{A}\psi|\hat{A}\psi\right>\left<\hat{B}\psi|\hat{B}\psi\right>} \\
&\ge \left|\left<\hat{A}\psi|\hat{B}\psi\right>\right| \quad \text{Cauchy-Schwartz}\\
&= \left|\left<\psi|\hat{A}\hat{B}\psi\right>\right| \\
\end{align} Δ A ^ Δ B ^ = ⟨ A ^ 2 ⟩ ⟨ B ^ 2 ⟩ = ⟨ A ^ ψ ∣ A ^ ψ ⟩ ⟨ B ^ ψ ∣ B ^ ψ ⟩ ≥ ⟨ A ^ ψ ∣ B ^ ψ ⟩ Cauchy-Schwartz = ⟨ ψ ∣ A ^ B ^ ψ ⟩
By similar argument, we have
Δ A ^ Δ B ^ ≥ ∣ < ψ ∣ B ^ A ^ ψ > ∣ \Delta \hat{A} \Delta \hat{B} \ge \left|\left<\psi|\hat{B}\hat{A}\psi\right>\right| Δ A ^ Δ B ^ ≥ ⟨ ψ ∣ B ^ A ^ ψ ⟩
Thus,
Δ A ^ Δ B ^ = 1 2 ( Δ A ^ Δ B ^ + Δ A ^ Δ B ^ ) ≥ 1 2 ( ∣ < ψ ∣ A ^ B ^ ψ > ∣ + ∣ < ψ ∣ B ^ A ^ ψ > ∣ ) ≥ 1 2 ∣ < ψ ∣ A ^ B ^ ψ > − < ψ ∣ B ^ A ^ ψ > ∣ = 1 2 ∣ < ψ ∣ [ A ^ , B ^ ] ψ > ∣ \begin{align}
\Delta \hat{A} \Delta \hat{B} &= \frac{1}{2}\left(\Delta \hat{A} \Delta \hat{B} + \Delta \hat{A} \Delta \hat{B}\right) \\
&\ge \frac{1}{2}\left(
\left|\left<\psi|\hat{A}\hat{B}\psi\right>\right| + \left|\left<\psi|\hat{B}\hat{A}\psi\right>\right|
\right)\\
&\ge \frac{1}{2}\left|
\left<\psi|\hat{A}\hat{B}\psi\right> - \left<\psi|\hat{B}\hat{A}\psi\right>
\right| \\
&= \frac{1}{2}\left|
\left<\psi|[\hat{A}, \hat{B}]\psi\right>
\right| \\
\end{align} Δ A ^ Δ B ^ = 2 1 ( Δ A ^ Δ B ^ + Δ A ^ Δ B ^ ) ≥ 2 1 ( ⟨ ψ ∣ A ^ B ^ ψ ⟩ + ⟨ ψ ∣ B ^ A ^ ψ ⟩ ) ≥ 2 1 ⟨ ψ ∣ A ^ B ^ ψ ⟩ − ⟨ ψ ∣ B ^ A ^ ψ ⟩ = 2 1 ⟨ ψ ∣ [ A ^ , B ^ ] ψ ⟩
Given the position operator x ^ \hat{x} x ^ and momentum operator p ^ \hat{p} p ^ , we have
Δ x ^ Δ p ^ ≥ ℏ 2 \Delta \hat{x} \Delta \hat{p} \ge \frac{\hbar}{2} Δ x ^ Δ p ^ ≥ 2 ℏ
A quantum harmonic oscillator is a system where the potential energy is given by
V ( x ) = 1 2 m ω 2 x 2 V(x) = \frac{1}{2}m\omega^2x^2 V ( x ) = 2 1 m ω 2 x 2
Given the Hamiltonian operator H ^ \hat{H} H ^ of the quantum harmonic oscillator, we can define the annihilation operator a ^ \hat{a} a ^ and creation operator a ^ † \hat{a}^\dagger a ^ † as
a ^ = u x ^ + v p ^ i a ^ † = u x ^ − v p ^ i \begin{align}
\hat{a} &= u\hat{x} + v\hat{p} i \\
\hat{a}^\dagger &= u\hat{x} - v\hat{p} i
\end{align} a ^ a ^ † = u x ^ + v p ^ i = u x ^ − v p ^ i
Thus, we have
a ^ a ^ † = u 2 x ^ 2 − v 2 p ^ 2 − u v x ^ p ^ i + u v p ^ x ^ i = u 2 x ^ 2 − v 2 p ^ 2 + u v ℏ \begin{align}
\hat{a}\hat{a}^\dagger &= u^2 \hat{x}^2 - v^2 \hat{p}^2
- uv \hat{x}\hat{p} i + uv \hat{p}\hat{x} i \\
&= u^2 \hat{x}^2 - v^2 \hat{p}^2 + uv\hbar \\
\end{align} a ^ a ^ † = u 2 x ^ 2 − v 2 p ^ 2 − uv x ^ p ^ i + uv p ^ x ^ i = u 2 x ^ 2 − v 2 p ^ 2 + uv ℏ
If we take:
u 2 = 1 2 ℏ m ω v 2 = 1 2 m ω ℏ \begin{align}
u^2 &= \frac{1}{2\hbar}m\omega \\
v^2 &= \frac{1}{2 m\omega\hbar} \\
\end{align} u 2 v 2 = 2ℏ 1 mω = 2 mω ℏ 1
Then, we have
a ^ a ^ † = 1 2 ℏ m ω x ^ 2 + 1 2 m ω ℏ p ^ 2 + 1 2 = 1 ℏ ω H ^ + 1 2 \begin{align}
\hat{a}\hat{a}^\dagger &= \frac{1}{2\hbar}m\omega \hat{x}^2 + \frac{1}{2 m\omega\hbar} \hat{p}^2 + \frac{1}{2} \\
&= \frac{1}{\hbar\omega}\hat{H} + \frac{1}{2}
\end{align} a ^ a ^ † = 2ℏ 1 mω x ^ 2 + 2 mω ℏ 1 p ^ 2 + 2 1 = ℏ ω 1 H ^ + 2 1
Given the annihilation operator a ^ \hat{a} a ^ and creation operator a ^ † \hat{a}^\dagger a ^ † of the quantum harmonic oscillator, we have
[ a ^ , a ^ † ] = 2 u v h = 1 [\hat{a}, \hat{a}^\dagger] = 2uvh = 1 [ a ^ , a ^ † ] = 2 uv h = 1
Given the annihilation operator a ^ \hat{a} a ^ and creation operator a ^ † \hat{a}^\dagger a ^ † of the quantum harmonic oscillator, we can define the number operator N ^ \hat{N} N ^ as
N ^ = a ^ † a ^ \hat{N} = \hat{a}^\dagger\hat{a} N ^ = a ^ † a ^
Thus, we have
H ^ = ℏ ω ( N ^ + 1 2 ) \begin{align}
\hat{H} = \hbar\omega\left(\hat{N} + \frac{1}{2}\right)
\end{align} H ^ = ℏ ω ( N ^ + 2 1 )
As the number operator N ^ \hat{N} N ^ is Hermitian, we can find an orthonormal basis { Ψ 0 ( x ) , Ψ 1 ( x ) , … } \{\Psi_0(x), \Psi_1(x), \ldots\} { Ψ 0 ( x ) , Ψ 1 ( x ) , … } such that
N ^ Ψ n ( x ) = E n Ψ n ( x ) \hat{N}\Psi_n(x) = E_n\Psi_n(x) N ^ Ψ n ( x ) = E n Ψ n ( x )
We next prove that E n E_n E n can only be non-negative integers.
Given Ψ n \Psi_n Ψ n an eigenvector of N ^ \hat{N} N ^ with eigenvalue E n E_n E n , we have
N ^ a ^ Ψ n = a ^ † a ^ 2 Ψ n = ( a ^ a ^ † − 1 ) a ^ Ψ n = a ^ N ^ − a ^ Ψ n = ( E n − 1 ) a ^ Ψ n \begin{align}
\hat{N}\hat{a}\Psi_n &= \hat{a}^\dagger\hat{a}^2\Psi_n \\
&= (\hat{a}\hat{a}^\dagger - 1)\hat{a}\Psi_n \\
&= \hat{a}\hat{N} - \hat{a}\Psi_n \\
&= (E_n-1)\hat{a}\Psi_n \\
\end{align} N ^ a ^ Ψ n = a ^ † a ^ 2 Ψ n = ( a ^ a ^ † − 1 ) a ^ Ψ n = a ^ N ^ − a ^ Ψ n = ( E n − 1 ) a ^ Ψ n
Thus, if a ^ Ψ n \hat{a}\Psi_n a ^ Ψ n is non-zero, then a ^ Ψ n \hat{a}\Psi_n a ^ Ψ n is also an eigenvector of N ^ \hat{N} N ^ with eigenvalue E n − 1 E_n-1 E n − 1 .
We next prove that a ^ Ψ n = 0 \hat{a}\Psi_n = 0 a ^ Ψ n = 0 if and only if E n = 0 E_n = 0 E n = 0 .
If E n = 0 E_n = 0 E n = 0 , then
< a ^ Ψ n ∣ a ^ Ψ n > = < Ψ n ∣ a ^ † a ^ Ψ n > = < Ψ n ∣ 0 > = 0 \begin{align}
\left<\hat{a}\Psi_n|\hat{a}\Psi_n\right> &= \left<\Psi_n|\hat{a}^\dagger\hat{a}\Psi_n\right> \\
&= \left<\Psi_n|0\right> \\
&= 0
\end{align} ⟨ a ^ Ψ n ∣ a ^ Ψ n ⟩ = ⟨ Ψ n ∣ a ^ † a ^ Ψ n ⟩ = ⟨ Ψ n ∣0 ⟩ = 0
Thus, a ^ Ψ n = 0 \hat{a}\Psi_n = 0 a ^ Ψ n = 0 .
If a ^ Ψ n = 0 \hat{a}\Psi_n = 0 a ^ Ψ n = 0 , then
< a ^ Ψ n ∣ a ^ Ψ n > = < Ψ n ∣ a ^ † a ^ Ψ n > = < Ψ n ∣ N ^ Ψ n > = E n < Ψ n ∣ Ψ n > = 0 \begin{align}
\left<\hat{a}\Psi_n|\hat{a}\Psi_n\right> &= \left<\Psi_n|\hat{a}^\dagger\hat{a}\Psi_n\right> \\
&= \left<\Psi_n|\hat{N}\Psi_n\right> \\
&= E_n\left<\Psi_n|\Psi_n\right> \\
&= 0
\end{align} ⟨ a ^ Ψ n ∣ a ^ Ψ n ⟩ = ⟨ Ψ n ∣ a ^ † a ^ Ψ n ⟩ = ⟨ Ψ n ∣ N ^ Ψ n ⟩ = E n ⟨ Ψ n ∣ Ψ n ⟩ = 0
Thus, E n = 0 E_n = 0 E n = 0 .
By proceeding the previous argument, we conclude that a ^ k Ψ n \hat{a}^k\Psi_n a ^ k Ψ n is an eigenvector of N ^ \hat{N} N ^ with eigenvalue E n − k E_n-k E n − k .
If E n En E n is not an integer, then k k k can be any positive integer as the annihilation process can be repeated indefinitely when E n − k E_n - k E n − k never hit zero.
And there exits k k k such that E n − k < 0 E_n - k < 0 E n − k < 0 , which is not possible,
as N ^ = a ^ † a ^ \hat{N} = \hat{a}^\dagger\hat{a} N ^ = a ^ † a ^ , and is positive semi-definite.
Thus, E n E_n E n is a non-negative integer.
By discussion in the previous section ,
we can find an normalised eigenstate Ψ 0 ( x ) \Psi_0(x) Ψ 0 ( x ) of the number operator N ^ \hat{N} N ^ with eigenvalue E 0 = 0 E_0 = 0 E 0 = 0 .
In this section, we prove that it is unique, up to a complex constant.
By previous discussion, we see that a ^ Ψ 0 = 0 \hat{a}\Psi_0 = 0 a ^ Ψ 0 = 0 .
u x ^ Ψ 0 + v p ^ i Ψ 0 = 0 u x ^ Ψ 0 + v ℏ d d x Ψ 0 = 0 \begin{align}
u\hat{x}\Psi_0 + v\hat{p} i \Psi_0 &= 0 \\
u\hat{x}\Psi_0 + v\hbar \frac{d}{dx}\Psi_0 &= 0 \\
\end{align} u x ^ Ψ 0 + v p ^ i Ψ 0 u x ^ Ψ 0 + v ℏ d x d Ψ 0 = 0 = 0
As the above equation is a first oder homogeneous ODE, and the solution is unique up to a complex constant.
Given the ground state Ψ 0 ( x ) \Psi_0(x) Ψ 0 ( x ) of the quantum harmonic oscillator, and eigenstate Ψ n ( x ) \Psi_n(x) Ψ n ( x ) of the number operator N ^ \hat{N} N ^ with eigenvalue E n = n E_n = n E n = n .
Then:
N ^ a ^ † Ψ n = ( E n + 1 ) a ^ † Ψ n \hat{N}\hat{a}^\dagger\Psi_n = (E_n+1) \hat{a}^\dagger\Psi_n N ^ a ^ † Ψ n = ( E n + 1 ) a ^ † Ψ n
and
< a ^ † Ψ n ∣ a ^ † Ψ n > = < Ψ n ∣ a ^ a ^ † Ψ n > = < Ψ n ∣ [ a ^ , a ^ † ] + a ^ † a ^ Ψ n > = < Ψ n ∣ Ψ n > + < Ψ n ∣ N ^ Ψ n > = E n + 1 \begin{align}
\left<\hat{a}^\dagger\Psi_n|\hat{a}^\dagger\Psi_n\right> &=
\left<\Psi_n|\hat{a}\hat{a}^\dagger\Psi_n\right> \\
&= \left<\Psi_n|[\hat{a},\hat{a}^\dagger]+\hat{a}^\dagger\hat{a}\Psi_n\right> \\
&= \left<\Psi_n|\Psi_n\right> + \left<\Psi_n|\hat{N}\Psi_n\right> \\
&= E_n + 1
\end{align} ⟨ a ^ † Ψ n ∣ a ^ † Ψ n ⟩ = ⟨ Ψ n ∣ a ^ a ^ † Ψ n ⟩ = ⟨ Ψ n ∣ [ a ^ , a ^ † ] + a ^ † a ^ Ψ n ⟩ = ⟨ Ψ n ∣ Ψ n ⟩ + ⟨ Ψ n ∣ N ^ Ψ n ⟩ = E n + 1
Thus, we could define the recursive relation:
Ψ n + 1 = a ^ † Ψ n E n + 1 \Psi_{n+1} = \frac{\hat{a}^\dagger\Psi_n}{\sqrt{E_n+1}} Ψ n + 1 = E n + 1 a ^ † Ψ n
Repeating the formula, we can get all the eigenstates of the number operator N ^ \hat{N} N ^ .
And could be defined by the ground state Ψ 0 ( x ) \Psi_0(x) Ψ 0 ( x ) .
Ψ n ( x ) = ( a ^ † ) n Ψ 0 ( x ) n ! \Psi_n(x) = \frac{(\hat{a}^\dagger)^n\Psi_0(x)}{\sqrt{n!}} Ψ n ( x ) = n ! ( a ^ † ) n Ψ 0 ( x )
As the creation operator a ^ † \hat{a}^\dagger a ^ † and annihilation operator a ^ \hat{a} a ^ are linear combinations of position operator x ^ \hat{x} x ^ and momentum operator p ^ \hat{p} p ^ ,
we can derive the position operator x ^ \hat{x} x ^ and momentum operator p ^ \hat{p} p ^ in terms of a ^ \hat{a} a ^ and a ^ † \hat{a}^\dagger a ^ † .
x ^ = 1 2 m ℏ ω ( a ^ + a ^ † ) p ^ = i 2 m ℏ ω ( a ^ − a ^ † ) \begin{align}
\hat{x} &= \frac{1}{\sqrt{2m\hbar\omega}}(\hat{a} + \hat{a}^\dagger) \\
\hat{p} &= \frac{i}{\sqrt{2m\hbar\omega}}(\hat{a} - \hat{a}^\dagger)
\end{align} x ^ p ^ = 2 m ℏ ω 1 ( a ^ + a ^ † ) = 2 m ℏ ω i ( a ^ − a ^ † )
Thus, we have
< Ψ n ∣ x ^ Ψ n > = 1 2 m ℏ ω < Ψ n ∣ ( a ^ + a ^ † ) Ψ n > = 0 \begin{align}
\left<\Psi_n|\hat{x}\Psi_n\right> &= \frac{1}{\sqrt{2m\hbar\omega}}\left<\Psi_n|(\hat{a} + \hat{a}^\dagger)\Psi_n\right> = 0 \\
\end{align} ⟨ Ψ n ∣ x ^ Ψ n ⟩ = 2 m ℏ ω 1 ⟨ Ψ n ∣ ( a ^ + a ^ † ) Ψ n ⟩ = 0
and
< Ψ n ∣ p ^ Ψ n > = i 2 m ℏ ω < Ψ n ∣ ( a ^ − a ^ † ) Ψ n > = 0 \begin{align}
\left<\Psi_n|\hat{p}\Psi_n\right> &= \frac{i}{\sqrt{2m\hbar\omega}}\left<\Psi_n|(\hat{a} - \hat{a}^\dagger)\Psi_n\right> = 0 \\
\end{align} ⟨ Ψ n ∣ p ^ Ψ n ⟩ = 2 m ℏ ω i ⟨ Ψ n ∣ ( a ^ − a ^ † ) Ψ n ⟩ = 0
Given an operator A ^ \hat{A} A ^ ,
which is probably time dependent,
then
d d t < A > = d d t < Ψ ∣ A ^ Ψ > = d d t ∫ − ∞ ∞ Ψ ∗ A ^ Ψ d x = ∫ − ∞ ∞ ( ∂ Ψ ∗ ∂ t A ^ Ψ + Ψ ∗ ∂ A ^ ∂ t Ψ + Ψ ∗ A ^ ∂ Ψ ∂ t ) d x = < Ψ ∣ ∂ A ^ ∂ t Ψ > + ∫ − ∞ ∞ ( ∂ Ψ ∗ ∂ t A ^ Ψ + Ψ ∗ A ^ ∂ Ψ ∂ t ) d x = < Ψ ∣ ∂ A ^ ∂ t Ψ > + < ∂ Ψ ∂ t ∣ A ^ Ψ > + < Ψ ∣ A ^ ∂ Ψ ∂ t > = < Ψ ∣ ∂ A ^ ∂ t Ψ > + < ( 1 i ℏ H ^ Ψ ) ∣ A ^ Ψ > + < Ψ ∣ A ^ ( 1 i ℏ H ^ Ψ ) > = < Ψ ∣ ∂ A ^ ∂ t Ψ > + 1 i ℏ [ − < Ψ ∣ H ^ A ^ Ψ > + < Ψ ∣ A ^ H ^ Ψ > ] = < Ψ ∣ ∂ A ^ ∂ t Ψ > + 1 i ℏ < [ A ^ , H ^ ] > \begin{align}
\frac{d}{dt}\left<A\right> &= \frac{d}{dt}\left<\Psi|\hat{A}\Psi\right> \\
&= \frac{d}{dt} \int_{-\infty}^{\infty} \Psi^*\hat{A}\Psi dx \\
&= \int_{-\infty}^{\infty} \left(\frac{\partial \Psi^*}{\partial t}\hat{A}\Psi + \Psi^*\frac{\partial \hat{A}}{\partial t}\Psi + \Psi^*\hat{A}\frac{\partial \Psi}{\partial t}\right) dx \\
&= \left<\Psi|\frac{\partial \hat{A}}{\partial t}\Psi\right>
+\int_{-\infty}^{\infty}
\left(\frac{\partial \Psi^*}{\partial t}\hat{A}\Psi + \Psi^*\hat{A}\frac{\partial \Psi}{\partial t}\right) dx \\
&= \left<\Psi|\frac{\partial \hat{A}}{\partial t}\Psi\right>
+ \left<\frac{\partial \Psi}{\partial t}|\hat{A}\Psi\right>
+ \left<\Psi|\hat{A}\frac{\partial \Psi}{\partial t}\right> \\
&= \left<\Psi|\frac{\partial \hat{A}}{\partial t}\Psi\right>
+ \left<\left(\frac{1}{i\hbar}\hat{H}\Psi\right)|\hat{A}\Psi\right>
+ \left<\Psi|\hat{A}\left(\frac{1}{i\hbar}\hat{H}\Psi\right)\right> \\
&= \left<\Psi|\frac{\partial \hat{A}}{\partial t}\Psi\right>
+ \frac{1}{i\hbar}\left[
- \left<\Psi|\hat{H}\hat{A}\Psi\right>
+ \left<\Psi|\hat{A}\hat{H}\Psi\right>
\right] \\
&= \left<\Psi|\frac{\partial \hat{A}}{\partial t}\Psi\right>
+ \frac{1}{i\hbar}
\left<[\hat{A},\hat{H}]\right>
\end{align} d t d ⟨ A ⟩ = d t d ⟨ Ψ∣ A ^ Ψ ⟩ = d t d ∫ − ∞ ∞ Ψ ∗ A ^ Ψ d x = ∫ − ∞ ∞ ( ∂ t ∂ Ψ ∗ A ^ Ψ + Ψ ∗ ∂ t ∂ A ^ Ψ + Ψ ∗ A ^ ∂ t ∂ Ψ ) d x = ⟨ Ψ∣ ∂ t ∂ A ^ Ψ ⟩ + ∫ − ∞ ∞ ( ∂ t ∂ Ψ ∗ A ^ Ψ + Ψ ∗ A ^ ∂ t ∂ Ψ ) d x = ⟨ Ψ∣ ∂ t ∂ A ^ Ψ ⟩ + ⟨ ∂ t ∂ Ψ ∣ A ^ Ψ ⟩ + ⟨ Ψ∣ A ^ ∂ t ∂ Ψ ⟩ = ⟨ Ψ∣ ∂ t ∂ A ^ Ψ ⟩ + ⟨ ( i ℏ 1 H ^ Ψ ) ∣ A ^ Ψ ⟩ + ⟨ Ψ∣ A ^ ( i ℏ 1 H ^ Ψ ) ⟩ = ⟨ Ψ∣ ∂ t ∂ A ^ Ψ ⟩ + i ℏ 1 [ − ⟨ Ψ∣ H ^ A ^ Ψ ⟩ + ⟨ Ψ∣ A ^ H ^ Ψ ⟩ ] = ⟨ Ψ∣ ∂ t ∂ A ^ Ψ ⟩ + i ℏ 1 ⟨ [ A ^ , H ^ ] ⟩