next up previous
Next: Rydberg Formula Up: Central Potentials Previous: Infinite Spherical Potential Well


Hydrogen Atom

A hydrogen atom consists of an electron, of charge $-e$ and mass $m_e$, and a proton, of charge $+e$ and mass $m_p$, moving in the Coulomb potential
\begin{displaymath}
V({\bf r}) = - \frac{e^2}{4\pi \epsilon_0 \vert{\bf r}\vert},
\end{displaymath} (661)

where ${\bf r}$ is the position vector of the electron with respect to the proton. Now, according to the analysis in Sect. 6.4, this two-body problem can be converted into an equivalent one-body problem. In the latter problem, a particle of mass
\begin{displaymath}
\mu = \frac{m_e m_p}{m_e+m_p}
\end{displaymath} (662)

moves in the central potential
\begin{displaymath}
V(r) = - \frac{e^2}{4\pi \epsilon_0 r}.
\end{displaymath} (663)

Note, however, that since $m_e/m_p\simeq 1/1836$ the difference between $m_e$ and $\mu$ is very small. Hence, in the following, we shall write neglect this difference entirely.

Writing the wavefunction in the usual form,

\begin{displaymath}
\psi(r,\theta,\phi) = R_{n,l}(r) Y_{l,m}(\theta,\phi),
\end{displaymath} (664)

it follows from Sect. 9.2 that the radial function $R_{n,l}(r)$ satisfies
\begin{displaymath}
-\frac{\hbar^2}{2 m_e}\left(\frac{d^2}{dr^2} + \frac{2}{r}\...
...} -\left(\frac{e^2}{4\pi \epsilon_0 r}+E
\right) R_{n,l}= 0.
\end{displaymath} (665)

Let $r = a z$, with
\begin{displaymath}
a = \sqrt{\frac{\hbar^2}{2 m_e (-E)}}=\sqrt{\frac{E_0}{E}} a_0,
\end{displaymath} (666)

where $E_0$ and $a_0$ are defined in Eqs. (678) and (679), respectively. Here, it is assumed that $E<0$, since we are only interested in bound-states of the hydrogen atom. The above differential equation transforms to
\begin{displaymath}
\left(\frac{d^2}{dz^2} + \frac{2}{z}\frac{d}{dz}-\frac{l (l+1)}{z^2}+ \frac{\zeta}{z}-1\right)
R_{n,l} = 0,
\end{displaymath} (667)

where
\begin{displaymath}
\zeta = \frac{2 m_e a e^2}{4\pi \epsilon_0 \hbar^2}=2 \sqrt{\frac{E_0}{E}}.
\end{displaymath} (668)

Suppose that $R_{n,l}(r) = Z(r/a) \exp(-r/a)/(r/a)$. It follows that
\begin{displaymath}
\left(\frac{d^2}{dz^2} -2 \frac{d}{dz} - \frac{l (l+1)}{z^2} + \frac{\zeta}{z}\right) Z = 0.
\end{displaymath} (669)

We now need to solve the above differential equation in the domain $z=0$ to $z= \infty$, subject to the constraint that $R_{n,l}(r)$ be square-integrable.

Let us look for a power-law solution of the form

\begin{displaymath}
Z(z) = \sum_k c_k z^k.
\end{displaymath} (670)

Substituting this solution into Eq. (669), we obtain
\begin{displaymath}
\sum_k c_k\left\{k (k-1) z^{k-2} - 2 k z^{k-1} - l (l+1) z^{k-2}
+ \zeta z^{k-1}\right\} = 0.
\end{displaymath} (671)

Equating the coefficients of $z^{k-2}$ gives the recursion relation
\begin{displaymath}
c_k \left[k (k-1)-l (l+1)\right] = c_{k-1} \left[2 (k-1) - \zeta\right].
\end{displaymath} (672)

Now, the power series (670) must terminate at small $k$, at some positive value of $k$, otherwise $Z(z)$ behaves unphysically as $z\rightarrow 0$ [i.e., it yields an $R_{n,l}(r)$ that is not square-integrable as $r\rightarrow 0$]. From the above recursion relation, this is only possible if $[k_{min} (k_{min}-1)-l (l+1)]=0$, where the first term in the series is $c_{k_{min}} z^{k_{min}}$. There are two possibilities: $k_{min}=-l$ or $k_{min}=l+1$. However, the former possibility predicts unphysical behaviour of $Z(z)$ at $z=0$. Thus, we conclude that $k_{min}=l+1$. Note that, since $R_{n,l}(r)\simeq Z(r/a)/(r/a)\simeq (r/a)^l$ at small $r$, there is a finite probability of finding the electron at the nucleus for an $l=0$ state, whereas there is zero probability of finding the electron at the nucleus for an $l>0$ state [i.e., $\vert\psi\vert^2=0$ at $r=0$, except when $l=0$].

For large values of $z$, the ratio of successive coefficients in the power series (670) is

\begin{displaymath}
\frac{c_k}{c_{k-1}} = \frac{2}{k},
\end{displaymath} (673)

according to Eq. (672). This is the same as the ratio of successive coefficients in the power series
\begin{displaymath}
\sum_k \frac{(2 z)^k}{k!},
\end{displaymath} (674)

which converges to $\exp(2 z)$. We conclude that $Z(z)\rightarrow
\exp(2 z)$ as $z\rightarrow\infty$. It thus follows that $R_{n,l}(r)\sim Z(r/a) \exp(-r/a)/(r/a)\rightarrow \exp(r/a)/(r/a)$ as $r\rightarrow\infty$. This does not correspond to physically acceptable behaviour of the wavefunction, since $\int\vert\psi\vert^2 dV$ must be finite. The only way in which we can avoid this unphysical behaviour is if the power series (670) terminates at some maximum value of $k$. According to the recursion relation (672), this is only possible if
\begin{displaymath}
\frac{\zeta}{2} = n,
\end{displaymath} (675)

where $n$ is an integer, and the last term in the series is $c_n z^n$. Since the first term in the series is $c_{l+1} z^{l+1}$, it follows that $n$ must be greater than $l$, otherwise there are no terms in the series at all. Finally, it is clear from Eqs. (666), (668), and (675) that
\begin{displaymath}
E = \frac{E_0}{n^2}
\end{displaymath} (676)

and
\begin{displaymath}
a = n a_0,
\end{displaymath} (677)

where
\begin{displaymath}
E_0 = -\frac{m_e e^4}{2 (4\pi \epsilon_0)^2 \hbar^2} = - \frac{e^2}{8\pi \epsilon_0 a_0}
= -13.6 {\rm eV},
\end{displaymath} (678)

and
\begin{displaymath}
a_0 = \frac{4\pi \epsilon_0 \hbar^2}{m_e e^2} = 5.3\times 10^{-11} {\rm m}.
\end{displaymath} (679)

Here, $E_0$ is the energy of so-called ground-state (or lowest energy state) of the hydrogen atom, and the length $a_0$ is known as the Bohr radius. Note that $\vert E_0\vert\sim \alpha^2 m_e c^2$, where $\alpha = e^2/ (4\pi \epsilon_0 \hbar c)\simeq 1/137$ is the dimensionless fine-structure constant. The fact that $\vert E_0\vert\ll m_e c^2$ is the ultimate justification for our non-relativistic treatment of the hydrogen atom.

We conclude that the wavefunction of a hydrogen atom takes the form

\begin{displaymath}
\psi_{n,l,m}(r,\theta,\phi) = R_{n,l}(r) Y_{l,m}(\theta,\phi).
\end{displaymath} (680)

Here, the $Y_{l,m}(\theta,\phi)$ are the spherical harmonics (see Sect 8.7), and $R_{n,l}(z=r/a)$ is the solution of
\begin{displaymath}
\left(\frac{1}{z^2} \frac{d}{dz} z^2 \frac{d}{dz}- \frac{l (l+1)}{z^2}
+ \frac{2 n}{z}-1\right) R_{n,l} = 0
\end{displaymath} (681)

which varies as $z^l$ at small $z$. Furthermore, the quantum numbers $n$, $l$, and $m$ can only take values which satisfy the inequality
\begin{displaymath}
\vert m\vert \leq l < n,
\end{displaymath} (682)

where $n$ is a positive integer, $l$ a non-negative integer, and $m$ an integer.

Now, we expect the stationary states of the hydrogen atom to be orthonormal: i.e.,

\begin{displaymath}
\int \psi^\ast_{n',l',m'} \psi_{n,l,m} dV = \delta_{nn'} \delta_{ll'} \delta_{mm'},
\end{displaymath} (683)

where $dV$ is a volume element, and the integral is over all space. Of course, $dV = r^2 dr d{\mit\Omega}$, where $d{\mit\Omega}$ is an element of solid angle. Moreover, we already know that the spherical harmonics are orthonormal [see Eq. (615)]: i.e.,
\begin{displaymath}
\oint Y_{l',m'}^{ \ast} Y_{l,m} d\Omega = \delta_{ll'} \delta_{mm'}.
\end{displaymath} (684)

It, thus, follows that the radial wavefunction satisfies the orthonormality constraint
\begin{displaymath}
\int_0^{\infty} R_{n',l}^\ast R_{n,l} r^2 dr = \delta_{nn'}.
\end{displaymath} (685)

The first few radial wavefunctions for the hydrogen atom are listed below:
$\displaystyle R_{1,0}(r)$ $\textstyle =$ $\displaystyle \frac{2}{a_0^{ 3/2}} \exp\left(-\frac{r}{a_0}\right),$ (686)
$\displaystyle R_{2,0}(r)$ $\textstyle =$ $\displaystyle \frac{2}{(2 a_0)^{3/2}}\left(1-\frac{r}{2 a_0}\right)
\exp\left(-\frac{r}{2 a_0}\right),$ (687)
$\displaystyle R_{2,1}(r)$ $\textstyle =$ $\displaystyle \frac{1}{\sqrt{3} (2 a_0)^{3/2}} \frac{r}{a_0} 
\exp\left(-\frac{r}{2 a_0}\right),$ (688)
$\displaystyle R_{3,0}(r)$ $\textstyle =$ $\displaystyle \frac{2}{(3 a_0)^{3/2}}\left(1-
\frac{2 r}{3 a_0} + \frac{2 r^2}{27 a_0^{ 2}}\right)\exp\left(-\frac{r}{3 a_0}\right),$ (689)
$\displaystyle R_{3,1}(r)$ $\textstyle =$ $\displaystyle \frac{4 \sqrt{2}}{9 (3 a_0)^{3/2}} \frac{r}{a_0}
\left(1-\frac{r}{6 a_0}\right) \exp\left(-\frac{r}{3 a_0}\right),$ (690)
$\displaystyle R_{3,2}(r)$ $\textstyle =$ $\displaystyle \frac{2 \sqrt{2}}{27 \sqrt{5} (3 a_0)^{3/2}}
\left(\frac{r}{a_0}\right)^2 \exp\left(-\frac{r}{3 a_0}\right).$ (691)

These functions are illustrated in Figs. 21 and 22.

Figure 21: The $a_0 r^2 \vert R_{n,l}(r)\vert^{ 2}$ plotted as a functions of $r/a_0$. The solid, short-dashed, and long-dashed curves correspond to $n,l=1,0$, and $2,0$, and $2,1$, respectively.
\begin{figure}
\epsfysize =3in
\centerline{\epsffile{Chapter09/fig02.eps}}
\end{figure}

Figure 22: The $a_0 r^2 \vert R_{n,l}(r)\vert^{ 2}$ plotted as a functions of $r/a_0$. The solid, short-dashed, and long-dashed curves correspond to $n,l=3,0$, and $3,1$, and $3,2$, respectively.
\begin{figure}
\epsfysize =3in
\centerline{\epsffile{Chapter09/fig03.eps}}
\end{figure}

Given the (properly normalized) hydrogen wavefunction (680), plus our interpretation of $\vert\psi\vert^2$ as a probability density, we can calculate

\begin{displaymath}
\langle r^k\rangle = \int_0^\infty r^{2+k} \vert R_{n,l}(r)\vert^{ 2} dr,
\end{displaymath} (692)

where the angle-brackets denote an expectation value. For instance, it can be demonstrated (after much tedious algebra) that
$\displaystyle \langle r^2\rangle$ $\textstyle =$ $\displaystyle \frac{a_0^{ 2} n^2}{2} [5 n^2+1-3 l (l+1)],$ (693)
$\displaystyle \langle r\rangle$ $\textstyle =$ $\displaystyle \frac{a_0}{2} [3 n^2-l (l+1)],$ (694)
$\displaystyle \left\langle \frac{1}{r}\right\rangle$ $\textstyle =$ $\displaystyle \frac{1}{n^2 a_0},$ (695)
$\displaystyle \left\langle\frac{1}{r^2}\right\rangle$ $\textstyle =$ $\displaystyle \frac{1}{(l+1/2) n^3 a_0^{ 2}},$ (696)
$\displaystyle \left\langle\frac{1}{r^3}\right\rangle$ $\textstyle =$ $\displaystyle \frac{1}{l (l+1/2) (l+1) n^3 a_0^{ 3}}.$ (697)

According to Eq. (676), the energy levels of the bound-states of a hydrogen atom only depend on the radial quantum number $n$. It turns out that this is a special property of a $1/r$ potential. For a general central potential, $V(r)$, the quantized energy levels of a bound-state depend on both $n$ and $l$ (see Sect. 9.3).

The fact that the energy levels of a hydrogen atom only depend on $n$, and not on $l$ and $m$, implies that the energy spectrum of a hydrogen atom is highly degenerate: i.e., there are many different states which possess the same energy. According to the inequality (682) (and the fact that $n$, $l$, and $m$ are integers), for a given value of $l$, there are $2 l+1$ different allowed values of $m$ (i.e., $-l,-l+1, \cdots, l-1, l$). Likewise, for a given value of $n$, there are $n$ different allowed values of $l$ (i.e., $0,1,\cdots, n-1$). Now, all states possessing the same value of $n$ have the same energy (i.e., they are degenerate). Hence, the total number of degenerate states corresponding to a given value of $n$ is

\begin{displaymath}
1 + 3 + 5 + \cdots +2 (n-1)+1 = n^2.
\end{displaymath} (698)

Thus, the ground-state ($n=1$) is not degenerate, the first excited state ($n=2$) is four-fold degenerate, the second excited state ($n=3$) is nine-fold degenerate, etc. [Actually, when we take into account the two spin states of an electron (see Sect. 10), the degeneracy of the $n$th energy level becomes $2 n^2$.]


next up previous
Next: Rydberg Formula Up: Central Potentials Previous: Infinite Spherical Potential Well
Richard Fitzpatrick 2010-07-20