The Schrödinger equation¶
Week 1, Lecture 3
So far we have been working with the time-dependent Schrödinger equation, which is a partial differential equation
In general, the potential in which the particle sits will depend on time, i.e. \(V=V(x,t)\). Let's now assume for a moment, that the potential is stable in time \(V(x,t)=V(x)\). While this seems very limiting at first, it's actually quite a common case. We can then separate the variables by introducing two new functions \(\Psi(x,t)=\psi(x)\cdot\phi(t)\) and write the TDSE as
We have no more partial derivates and we can re-arrange the equation such that the time dependent parts are on the left and the position dependent on the right
As the two sides depend on different variables and are equal to one another, the only possible solution is a constant, which we call \(E\). The position dependent part is called the time-independent Schrödinger equation (TISE):
while the time-dependent part simplifies to
We ended up with two ordinary differential equations instead of one partial. As a next step, we will try to find a general solution:
-
For the time-dependent part we guess the simple, general solution to be related to \(e^{ax}\), as it takes the right form \(\frac{\mathrm{d}}{\mathrm{d}x}e^{ax}=a\cdot e^{ax}\). Hence the general solution is \(\phi(t)=e^{-iEt/\hbar}\).
-
In order to solve the TISE we will need to specify the potential \(V(x)\). This is exactly what we will do for the rest of the course: choose a specific \(V(x)\) and then solve the TISE.
Note
In order to make the distinction between variables and operators clear, we will from now on use "hats" on operators: \(\hat{x}\), \(\hat{p}\), etc.
The time-independent Schrödinger equation is an eigenvalue differential equation, with the set of solutions \(\psi_n(x)\) and eigenvalues \(E_n\). These seperable solutions have some interesting properties:
-
They are stationary states, which means "nothing" except for the wavefunction changes in time. While \(\Psi(x,t)=\psi_n(x)e^{-iE_nt/\hbar}\) is time dependent, \(|\Psi(x,t)|^2\) is not. This is easy to see:
\[|\Psi(x,t)|^2=\psi_n^*e^{iE_nt/\hbar}\psi_ne^{-iE_nt/\hbar}=|\psi_n(x)|^2\] -
The solutions have no uncertainty in their total energy! From classical mechanics you are familiar with the energy \(E=\frac{1}{2}mv^2+V=\frac{p^2}{2m}+V(x)=H(x,p)\), which is also called the Hamiltonian. In quantum mechanics we have something similar: remember the operator for momentum \(\hat{p}=\frac{\hbar}{i}\frac{\mathrm{d}}{\mathrm{d}x}\), which we can use to write the total energy as
\[\hat{E}=\frac{1}{2m}\left(\frac{\hbar}{i}\frac{\mathrm{d}}{\mathrm{d}x}\right)\left(\frac{\hbar}{i}\frac{\mathrm{d}}{\mathrm{d}x}\right)+V(x)=\frac{-\hbar^2}{2m}\frac{\mathrm{d}^2}{\mathrm{d}x^2}+V(x)=\hat{H}\]Here \(\hat{H}\) is the Hamiltonian in QM and the TISE then simplifies to \(\hat{H}\psi_n=E_n\psi_n\). Now, in order to demonstrate that the solutions \(\psi_n\) have no uncertainty in energy, we need to calculate the standard deviation \(\sigma_E\) of the energy for \(\psi_n\): \(\sigma_E=\sqrt{\langle E^2\rangle-\langle E\rangle^2}\).
\[\langle E\rangle=\int_{-\infty}^{+\infty} \psi_n^*(x)e^{iE_nt/\hbar}\hat{H}\psi_ne^{-iE_nt/\hbar} \,\mathrm{d}x = \int_{-\infty}^{+\infty} \psi_n^*(x)E_n\psi_n(x) \,\mathrm{d}x = \]\[=E_n \int_{-\infty}^{+\infty} \psi_n^*(x)\psi_n(x) \,\mathrm{d}x = E_n\]\[\langle E^2\rangle=\int_{-\infty}^{+\infty} \psi_n^*(x)\hat{H}^2\psi_n \,\mathrm{d}x=\int_{-\infty}^{+\infty} \psi_n^*(x)E_n^2\psi_n \,\mathrm{d}x=E_n^2\]\[\sigma_E^2=\langle E^2\rangle-\langle E\rangle^2=E_n^2-E_n^2=0\]Hence \(\psi_n\) have definite energy!
-
The expectation value of any operator \(\hat{Q}\) is independent of time.
\[\langle Q\rangle=\int_{-\infty}^{+\infty} \Psi*(x,t)\hat{Q}\left(x,\frac{\mathrm{d}}{\mathrm{d}x}\right)\Psi(x,t)\,\mathrm{d}x =\]\[=\int_{-\infty}^{+\infty} \psi_n^*(x)e^{iE_nt/\hbar}\hat{Q}\left(x,\frac{\mathrm{d}}{\mathrm{d}x}\right)\psi_n(x)e^{-iE_nt/\hbar}\,\mathrm{d}x =\]\[=\int_{-\infty}^{+\infty} \psi_n^*(x)\hat{Q}\left(x,\frac{\mathrm{d}}{\mathrm{d}x}\right)\psi_n(x)\,\mathrm{d}x\] -
The solutions \(\psi_n\) form a complete and orthonormal set:
- We can perform a Fourier decomposition of any arbitrary function: \(f(x)=\sum_n c_n \psi_n(x)\)
- \(\psi_n\) are normalized: \(\int_{-\infty}^{+\infty} \psi_n^*\psi_n \,\mathrm{d}x = 1\)
- \(\int_{-\infty}^{+\infty} \psi_n^*\psi_m \,\mathrm{d}x = \delta_{n,m}\), where \(\delta_{n,m}\) is the Kronecker delta.
Note
Recipe for solving the TDSE with \(V=V(x)\)
-
Step 1: expand \(\Psi(x,0)=\sum_n c_n\psi_n(x)\) \(\int_{-\infty}^{+\infty} \psi_m^*\Psi(x,0) \,\mathrm{d}x = \int_{-\infty}^{+\infty} \psi_m^*\sum_n c_n\psi_n(x) \,\mathrm{d}x = \sum_n c_n\int_{-\infty}^{+\infty} \psi_m^*\psi_n(x) \,\mathrm{d}x = \sum_n c_n \delta_{m,n} = c_m\) \(\rightarrow c_n=\int_{-\infty}^{+\infty}\psi_n^*\Psi(x,0)\,\mathrm{d}x\)
-
Step 2: add the time-dependence of \(\psi_n\). If \(\Psi_n(x,t=0)=\psi_n(x) \rightarrow \Psi_n(x,t)=e^{-iE_nt/\hbar}\psi_n(x)\)
-
Step 3: use the superposition principle \(\Psi(x,t)=\sum_n c_n e^{-iE_nt/\hbar}\psi_n(x)\)
Example:
A particle is prepared in a superposition state (i.e. the linear combination of stationary states): \(\Psi(x,0)=c_1\psi_1+c_2\psi_2\). We know that \(\psi_1\) and \(\psi_2\) are stationary states with energies \(E_1\) and \(E_2\), respectively.
-
What is \(\Psi(x,t)\)? \(\Psi(x,t)=c_1e^{-iE_1t/\hbar}\psi_1+c_2e^{-iE_2t/\hbar}\psi_2\)
-
What is \(|\Psi(x,t)|^2\)?
For simplicity, we have assumed \(c_1\), \(c_2\) and \(\psi_1\), \(\psi_2\) to be real.
The \(|c_n|^2\) gives the probability of measuring energy \(E_n\). This is immediately clear if \(\Psi=\psi_1\) and for a superposition state \(\Psi=c_1\psi_1+c_2\psi_2\) it implies that \(\sum_n|c_n|^2=1\), i.e. normalization. It turns out that for any \(\Psi=\sum c_n\psi_n\), the expectation value \(\langle E\rangle\) is independent of time. This is of course the familiar conservation of energy! Please note that a measurement does not preserve energy.