Differences
This shows you the differences between two versions of the page.
Next revision | Previous revision | ||
the_schroedinger_equation [2021/03/15 21:08] – created admin | the_schroedinger_equation [2021/03/17 21:25] (current) – [3.iii.3 The Continuity Equation] admin | ||
---|---|---|---|
Line 10: | Line 10: | ||
====== 3.iii.1 Argument 1: Plane Wave Solutions ====== | ====== 3.iii.1 Argument 1: Plane Wave Solutions ====== | ||
- | Consider a free particle in one-dimension. | + | Consider a free particle in one-dimension. |
\[\frac{\partial^2 \psi}{\partial x^2} = \frac{1}{v^2}\frac{\partial^2 \psi}{\partial t^2},\] | \[\frac{\partial^2 \psi}{\partial x^2} = \frac{1}{v^2}\frac{\partial^2 \psi}{\partial t^2},\] | ||
where $v$ is the wave speed. | where $v$ is the wave speed. | ||
Line 16: | Line 16: | ||
Substituting $\psi(x,t) = Ae^{i(kx-\omega t)}$ into the wave equation gives | Substituting $\psi(x,t) = Ae^{i(kx-\omega t)}$ into the wave equation gives | ||
\[-Ak^2 e^{i(kx-\omega t)} = -\frac{A}{v^2}\omega^2 e^{i(kx-\omega t)}.\] | \[-Ak^2 e^{i(kx-\omega t)} = -\frac{A}{v^2}\omega^2 e^{i(kx-\omega t)}.\] | ||
- | Cancelling the common terms on both sides gives | + | Cancelling the common terms on both sides and rearranging |
\[\omega^2 = k^2v^2,\] | \[\omega^2 = k^2v^2,\] | ||
or | or | ||
Line 26: | Line 26: | ||
which implies that the classical wave equation cannot be the correct equation of motion for such a particle. | which implies that the classical wave equation cannot be the correct equation of motion for such a particle. | ||
- | How should the wave equation be modified to get the correct energy-momentum equation? | + | How should the wave equation be modified to get the correct energy-momentum equation? |
\[\frac{\partial \psi}{\partial t} = B \frac{\partial^2 \psi}{\partial x^2},\] | \[\frac{\partial \psi}{\partial t} = B \frac{\partial^2 \psi}{\partial x^2},\] | ||
where $B$ is a constant to be determined. | where $B$ is a constant to be determined. | ||
- | In an in class activity, you will show that to get the relation $E = p^2/2m$$, we need B = \frac{i\hbar}{2m}$. | + | In an in class activity, you will show that to get the relation $E = p^2/2m$, we need $B = \frac{i\hbar}{2m}$. |
\[\frac{\partial \psi}{\partial t} = \frac{i\hbar}{2m} \frac{\partial^2 \psi}{\partial x^2},\] | \[\frac{\partial \psi}{\partial t} = \frac{i\hbar}{2m} \frac{\partial^2 \psi}{\partial x^2},\] | ||
and multiplying both sides by $i\hbar$ gives | and multiplying both sides by $i\hbar$ gives | ||
Line 41: | Line 41: | ||
The same argument works in three dimensions, which gives the same equation with | The same argument works in three dimensions, which gives the same equation with | ||
- | \[\hat{H} = \frac{\har{p}^2}{2m} \rightarrow -\frac{\hbar^2}{2m} \nabla^2.\] | + | \[\hat{H} = \frac{\hat{p}^2}{2m} \rightarrow -\frac{\hbar^2}{2m} \nabla^2.\] |
- | Of course, this argument only works for a free particle. | + | Of course, this argument only works for a free particle. |
====== 3.iii.2 Argument 2: Conservation of Probability ====== | ====== 3.iii.2 Argument 2: Conservation of Probability ====== | ||
- | The first step of this argument is to assume the superposition principle. | + | The first step of this argument is to assume the superposition principle. |
+ | * $a\ket{\psi_1(t_0)} + b\ket{\psi_1(t_0)}$ is another possible solution at time $t_0$ and, | ||
+ | * it evolves in time to $a\ket{\psi_1(t)} + b\ket{\psi_1(t)}$ at time $t$. | ||
This means that the time evolution of the system between $t_0$ and $t$ is described by a linear operator $\hat{U}(t, | This means that the time evolution of the system between $t_0$ and $t$ is described by a linear operator $\hat{U}(t, | ||
- | \[\ket{\psi(t} = \hat{U}(t, | + | \[\ket{\psi(t)} = \hat{U}(t, |
The second thing to note is that the dynamics ought to preserve the norm of the state. | The second thing to note is that the dynamics ought to preserve the norm of the state. | ||
Line 58: | Line 60: | ||
The Born rule says that the probability of getting the outcome $a_n$ when measuring $\hat{A}$ at time $t_0$ is | The Born rule says that the probability of getting the outcome $a_n$ when measuring $\hat{A}$ at time $t_0$ is | ||
\[p(a_n, | \[p(a_n, | ||
- | This implies, the probability of obtaining any outcome at all, which should be $=1$, is | + | This implies, the probability of getting // |
\[\sum_{n}p(a_n, | \[\sum_{n}p(a_n, | ||
where we have used the Dirac Notaty to remove the identity $\hat{I} = \sum_n \proj{a_n}$. | where we have used the Dirac Notaty to remove the identity $\hat{I} = \sum_n \proj{a_n}$. | ||
Line 64: | Line 66: | ||
Now consider the situation at time $t$. The probability of getting the outcome $a_n$ when $\hat{A}$ is measured at time $t$ is | Now consider the situation at time $t$. The probability of getting the outcome $a_n$ when $\hat{A}$ is measured at time $t$ is | ||
\begin{align*} | \begin{align*} | ||
- | p(a_n,t) & = \Qprob{a_n}{\psi(t)}) \\ | + | p(a_n,t) & = \QProb{a_n}{\psi(t)}) \\ |
- | & \Abs{\sand{a_n}{\hat{U}(t, | + | & = \Abs{\sand{a_n}{\hat{U}(t, |
- | & = \sand{\psi(t_0)}{\hat{U}(t, | + | & = \sand{\psi(t_0)}{\hat{U}(t, |
\end{align*} | \end{align*} | ||
so, applying the Dirac Notaty again, the probability of obtaining any outcome at all is | so, applying the Dirac Notaty again, the probability of obtaining any outcome at all is | ||
- | \[sum_n p(a_n,t) = \sand{\psi(t_0)}{\hat{U}^{\dagger}(t, | + | \[\sum_n p(a_n,t) = \sand{\psi(t_0)}{\hat{U}^{\dagger}(t, |
- | We want this to equal $1$ as well, so, in particular, | + | We want this to equal $1$ as well, so, in particular, |
- | \[\sand{\psi(t_0)}{\hat{U}^{\dagger}(t, | + | \[\sand{\psi(t_0)}{\hat{U}^{\dagger}(t, |
which we can also write as | which we can also write as | ||
- | \[\sand{\psi(t_0)}{\hat{U}^{\dagger}(t, | + | \[\sand{\psi(t_0)}{\hat{U}^{\dagger}(t, |
- | Now, if $\ket{\psi(t_0)}$ can be any vector in the Hilbert space then we have that | + | Now, suppose that the initial state $\ket{\psi(t_0)}$ can be any vector in the Hilbert space. Then, this implies |
- | \[\sand{\psi}{\hat{U}^{\dagger}(t, | + | \[\sand{\psi}{\hat{U}^{\dagger}(t, |
for all vectors $\ket{\psi}$. | for all vectors $\ket{\psi}$. | ||
Now, we also want that $\hat{U}(t_0, | Now, we also want that $\hat{U}(t_0, | ||
- | \[\hat{U}(t, | + | \[\hat{U}(t, |
In an in class activity, you will show that the unitarity of $\hat{U}(t, | In an in class activity, you will show that the unitarity of $\hat{U}(t, | ||
Line 95: | Line 97: | ||
\[i\hbar\frac{\partial \ket{\psi}}{\partial t} = \hat{H}\ket{\psi}.\] | \[i\hbar\frac{\partial \ket{\psi}}{\partial t} = \hat{H}\ket{\psi}.\] | ||
- | This shows that the Schrödinger equation must be correct for //some// Hermitian operator $\hat{H}$, but not that $\hat{H}$ must be the Hamiltonian (the operator representing the energy). More sophisticated arguments can show that the generator of time translations, | + | This shows that the Schrödinger equation must be correct for //some// Hermitian operator $\hat{H}$, but not that $\hat{H}$ must be the Hamiltonian (the operator representing the energy). |
====== 3.iii.3 The Continuity Equation ====== | ====== 3.iii.3 The Continuity Equation ====== | ||
The previous argument shows that the total probability is conserved in time by Schrödinger evolution, i.e. | The previous argument shows that the total probability is conserved in time by Schrödinger evolution, i.e. | ||
- | \[\braket{\psi(t)}{\psi(t)} = \braket{\psi(0)}{\psi(0)}, | + | \[\braket{\psi(t)}{\psi(t)} = \braket{\psi(t_0)}{\psi(t_0)}, |
In fact, we can derive a more detailed result describing the //local// conservation of probability. | In fact, we can derive a more detailed result describing the //local// conservation of probability. | ||
- | The basic idea should be familiar if you have studied continuity equations in fluid dynamics or electromagnetism. | + | The basic idea should be familiar if you have studied continuity equations in fluid dynamics or electromagnetism. |
\[\frac{\partial \rho(\vec{r}, | \[\frac{\partial \rho(\vec{r}, | ||
where $\rho(\vec{r}, | where $\rho(\vec{r}, | ||
Line 110: | Line 112: | ||
In quantum mechanics, we are going to derive an equation like this for the // | In quantum mechanics, we are going to derive an equation like this for the // | ||
- | It is worth stopping to consider the meaning of this before we proceed. | + | It is worth stopping to consider the meaning of this before we proceed. |
Having understood this slightly weird concept, we know that $\rho(\vec{r}, | Having understood this slightly weird concept, we know that $\rho(\vec{r}, | ||
+ | |||
+ | To do this, we start with the Schrödinger equation written out in the position representation. | ||
+ | \[i\hbar \frac{\partial \psi(\vec{r}, | ||
+ | We now multiply this equation by the complex conjugate of the wavefunction $\psi^*(\vec{r}, | ||
+ | \begin{equation} | ||
+ | \label{eq1} | ||
+ | i\hbar \psi^*(\vec{r}, | ||
+ | \end{equation} | ||
+ | |||
+ | Next, we take the complex conjugate of the Schrödinger equation, which is | ||
+ | \[-i\hbar \frac{\partial \psi^*(\vec{r}, | ||
+ | and we multiply this equation by the wavefunction $\psi(\vec{r}, | ||
+ | \begin{equation} | ||
+ | \label{eq2} | ||
+ | -i\hbar \frac{\partial \psi^*(\vec{r}, | ||
+ | \end{equation} | ||
+ | |||
+ | The next step is to subtract equation \eqref{eq2} from equation \eqref{eq1}, | ||
+ | \[i\hbar \left [ \psi^*(\vec{r}, | ||
+ | Before doing anything else, let's just divide both sides of this equation by $i\hbar$, which gives | ||
+ | \begin{equation} | ||
+ | \label{eq3} | ||
+ | \psi^*(\vec{r}, | ||
+ | \end{equation} | ||
+ | |||
+ | Using the product rule for derivatives, | ||
+ | \[\frac{\partial \rho(\vec{r}, | ||
+ | so the left hand side of equation (\ref{eq3}) is just $\frac{\partial \rho(\vec{r}, | ||
+ | |||
+ | Now, we define the // | ||
+ | \[\boxed{\vec{J}(\vec{r}, | ||
+ | Again, by the product rule for derivatives, | ||
+ | \begin{align*} | ||
+ | \vec{\nabla} \cdot \vec{J}(\vec{r}, | ||
+ | & = \frac{i\hbar}{2m} \left [ \psi(\vec{r}, | ||
+ | \end{align*} | ||
+ | In other words, the right hand side of equation \eqref{eq3} is $-\vec{\nabla}\cdot \vec{J}(\vec{r}, | ||
+ | \[\frac{\partial \rho(\vec{r}, | ||
+ | for the probability density $\rho(\vec{r}, | ||
+ | \[\vec{J}(\vec{r}, | ||
+ | |||
+ | If you think about it, what we have just done is pretty weird. | ||
+ | |||
+ | Instead, what we have done here is start by assuming that there is some continuity equation for $\rho(\vec{r}, | ||
+ | \[\vec{J}' | ||
+ | where $\vec{F}(\vec{r}, | ||
+ | |||
+ | Nonetheless, | ||
+ | |||
+ | ====== 3.iii.4 Formal Solution of the Schrödinger Equation ====== | ||
+ | |||
+ | We saw in section 3.iii.2 that the time-evolution of a quantum system from time $t_0$ to time $t > t_0$ is described by a unitary operator $\hat{U}(t, | ||
+ | \[\ket{\psi(t)} = \hat{U}(t, | ||
+ | and such that $\hat{U}(t_0, | ||
+ | |||
+ | Using the Schrödinger equation, we can find an equation for $\hat{U}(t, | ||
+ | \[i\hbar \partial{\ket{\psi(t)}}{\partial t} = \hat{H} \ket{\psi(t)}, | ||
+ | but this is equivalent to | ||
+ | \[i\hbar \partial{\hat{U}(t, | ||
+ | Now, $\ket{\psi(t_0)}$ does not depend on $t$ and this equation has to be true for all possible initial states $\ket{\psi(t_0)}$, | ||
+ | \[i\hbar \partial{\hat{U}(t, | ||
+ | Dividing both sides by $i\hbar$ gives | ||
+ | \begin{equation} | ||
+ | \label{eq4} | ||
+ | \partial{\hat{U}(t, | ||
+ | \end{equation} | ||
+ | |||
+ | The unitary operator $\hat{U}(t, | ||
+ | is independent of time, i.e. since the kinetic energy term is independent of time, this happens when the potential $V(\vec{r}, | ||
+ | \[\boxed{\hat{U}(t, | ||
+ | This is known as the // | ||
+ | |||
+ | To understand why $\hat{U}(t, | ||
+ | \[\hat{B}(t) = e^{\alpha\hat{A}t}, | ||
+ | where $\alpha$ is a constant and $\hat{A}$ is independent of time, then | ||
+ | \[\hat{B}(t) = \sum_{n=0}^{\infty} \frac{\alpha^n \hat{A}^n}{n!}t^n, | ||
+ | and hence | ||
+ | \begin{align*} | ||
+ | \frac{\partial \hat{B}}\partial{t} & = \sum_{n=0}^{\infty} \frac{n\alpha^n \hat{A}^n}{n!}t^{n-1}. | ||
+ | \end{align*} | ||
+ | Due to the factor of $n$, the $n=0$ term is zero, so we have | ||
+ | \begin{align*} | ||
+ | \frac{\partial \hat{B}}\partial{t} & = \sum_{n=1}^{\infty} \frac{n\alpha^n \hat{A}^n}{n!}t^{n-1} \\ | ||
+ | & = \sum_{n=1}^{\infty} \frac{\alpha^n \hat{A}^n}{(n-1)!}t^{n-1} \\ | ||
+ | & = \alpha \hat{A}^n \left ( \sum_{n=1}^{\infty} \frac{\alpha^{n-1} \hat{A}^{n-1}}{(n-1)!}t^{n-1}\right ) \\ | ||
+ | & = \alpha \hat{A} \left ( \sum_{n=0}^{\infty} \frac{\alpha^{n} \hat{A}^{n}}{n!}t^{n}\right ) \\ | ||
+ | & = \alpha \hat{A} e^{a\hat{A}t}. | ||
+ | \end{align*} | ||
+ | Using this result, it is easily checked that $\hat{U}(t-t_0) = \hat{C}e^{-i(t-t_0)\hat{H}/ | ||
+ | \[\hat{I} = \hat{C}e^{0} = \hat{C}\hat{I} = \hat{C},\] | ||
+ | so we must have $\hat{C}=\hat{I}$ and so $\hat{U}(t-t_0) = e^{-i(t-t_0)\hat{H}/ | ||
+ | |||
+ | ====== 3.iii.5 From the Formal Solution to Useful Solutions ====== | ||
+ | |||
+ | At this point, we have shown that the solution to the Schrödinger equation is | ||
+ | \[\ket{\psi(t)} = e^{-i(t-t_0)\hat{H}/ | ||
+ | However, I do not recommend trying to solve practical problems by writing down $e^{-i(t-t_0)\hat{H}/ | ||
+ | |||
+ | Recall that, since $\hat{H}$ is a Hermitian operator, its eigenvectors form a complete orthonormal basis for the Hilbert space. | ||
+ | |||
+ | Now, recall that if $\ket{a}$ is an eigenvector of $\hat{A}$ with eigenvalue $a$ then $f(\hat{A})\ket{a} = f(a)\ket{a}$, | ||
+ | \[e^{i(t - t_0)\hat{H}/ | ||
+ | This equation says that if the system starts in an energy eigenstate $\ket{E_n, | ||
+ | |||
+ | The system need not start out in an energy eigenstate, but, since the states $\ket{E_n, | ||
+ | \[\ket{\psi(t_0)} = \sum_{n,m} c_{n, | ||
+ | where the components $c_{n, | ||
+ | \[\ket{\psi(t)} = \sum_{n,m} c_{n,m}(t) \ket{E_n, | ||
+ | where the new components $c_{n,m}(t) = \braket{E_n, | ||
+ | \begin{align*} | ||
+ | | ||
+ | & = e^{-i(t-t_0)\hat{H}/ | ||
+ | & = \sum_{n,m} c_{n, | ||
+ | & = \sum_{n,m} c_{n, | ||
+ | \end{align*} | ||
+ | from which we deduce that | ||
+ | \[\boxed{c_{n, | ||
+ | |||
+ | In other words, to solve the Schrödinger equation for an arbitrary initial state, you have to do the following. | ||
+ | |||
+ | * Solve the eigenvalue equation $\hat{H}\ket{E_n, | ||
+ | * Write your initial state as a superposition of energy eigenstates | ||
+ | \[\ket{\psi(t_0)} = \sum_{n,m} c_{n, | ||
+ | where $c_{n, | ||
+ | * At some later time $t$, this state will evolve to | ||
+ | \[\ket{\psi(t)} = \sum_{n,m} c_{n, | ||
+ | | ||
+ | Let's illustrate this method with a simple example. | ||
+ | \[\hat{H} = E \left ( \ketbra{\phi_1}{\phi_2} + \ketbra{\phi_2}{\phi_1}\right ),\] | ||
+ | where $E$ is a constant with the dimensions of energy. | ||
+ | |||
+ | It will be convenient to do the calculation in matrix form in the basis $\ket{\phi_1}, | ||
+ | \[H = E \left ( \begin{array}{cc} 0 & 1 \\ 1 & 0 \end{array}\right ),\] | ||
+ | and the initial state is represented by the column vector | ||
+ | \[\vec{\psi}(t_0) = \left ( \begin{array}{c} 1 \\ 0 \right ).\] | ||
+ | |||
+ | The eigenvalues of $H$ are of the form $\lambda E$, where $\lambda$ is an eigenvalue of | ||
+ | \[\left ( \begin{array}{cc} 0 & 1 \\ 1 & 0 \end{array}\right ),\] | ||
+ | so let's find the eigenvalues and eigenvectors of this matrix. | ||
+ | |||
+ | The characteristic equation is | ||
+ | \[(-\lambda)^2 - 1 = 0,\] | ||
+ | so $\lambda = \pm 1$. The eigenvalues of $\hat{H}$ are therefore $\pm E$. Since we have two distinct eigenvalues, | ||
+ | |||
+ | For $\lambda = 1$, the eigenvector is found by solving | ||
+ | \[\left ( \begin{array}{cc} -1 & 1 \\ 1 & -1 \end{array}\right ) \left ( \begin{array}{c} x \\ y \end{array} \right ) = \left ( \begin{array}{c} 0 \\ 0 \end{array} \right ),\] | ||
+ | which has solution $y=x$. | ||
+ | \[\vec{\chi}_1 = C\left ( \begin{array}{c} 1 \\ 1 \end{array} \right ),\] | ||
+ | for an arbitrary constant $C$. But, as I keep telling you and someone will nevertheless fail to do on a homework or exam, we should always work with normalized eigenvectors, | ||
+ | \[\vec{chi}_1 = \frac{1}{\sqrt{2}}\left ( \begin{array}{c} 1 \\ 1 \end{array} \right ).\] | ||
+ | |||
+ | For For $\lambda = -1$, the eigenvector is found by solving | ||
+ | \[\left ( \begin{array}{cc} 1 & 1 \\ 1 & 1 \end{array}\right ) \left ( \begin{array}{c} x \\ y \end{array} \right ) = \left ( \begin{array}{c} 0 \\ 0 \end{array} \right ),\] | ||
+ | which has solution $y=-x$. | ||
+ | \[\vec{\chi}_2 = \sqrt{1}{\sqrt{2}}\left ( \begin{array}{c} 1 \\ -1 \end{array} \right ),\] | ||
+ | where I have normalized the vector already. | ||
+ | |||
+ | Now, we have to express our initial state as a superposition of $\vec{\chi}_1$ and $\vec{\chi_2}$ |