0810.1019/A4.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \chapter{Spectroscopy and spectra}\label{c.spec}
4: 
5: This final chapter applies the Lie theoretic structure to the analysis
6: of quantum spectra. After a short history of some aspects of 
7: spectroscopy, we look at the spectrum of bound systems of particles.
8: We show how to obtain from a measured spectrum the spectrum of the
9: associated Hamiltonian, and discuss qualitative results on vibrations 
10: (giving discrete spectra) and chemical reactions (giving continuous 
11: spectra) that come from the consideration of simple systems and the 
12: consideration of approximate symmetries. The latter are shown to 
13: result in a clustering of spectral values.
14: 
15: The structure of the clusters is determined by how the irreducible 
16: representations of a dynamical Lie algebra split when the algebra is 
17: reduced to a subalgebra of generating symmetries. 
18: The clustering can also occur in a hierarchical fashion with fine 
19: splitting and hyperfine splitting, corresponding to a chain of 
20: subgroups.
21: As an example, we discuss the spectrum of the hydrogen atom.
22: 
23: 
24: 
25: 
26: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
27: \section{Introduction and historical background}
28: 
29: In this chapter we show some features of spectra and
30: spectroscopy. In the preceding chapters we discussed properties of
31: systems. The Hamiltonian of a system has a spectrum consisting of the
32: eigenvalues, but in practice we don't see this spectrum, but the 
33: energy differences. One perturbs the system by shining
34: light on it for example and then observes some response. The responses
35: give rise to the observed spectrum, the study of which is spectroscopy.
36: 
37: To study the structure of molecules and atoms, we often rely
38: on destructive methods. The destructive nature of the experiments in
39: chemistry was taken as a primitive distinction between chemistry and
40: physics. Nowadays the situation is different. In
41: high-energy physics, one also shoots particles at each other such
42: that the original particles are destroyed and energy is converted into
43: the creation of other particles. On the other side, in chemistry new
44: laser-techniques are used where molecules are kept intact,
45: and information about the structure of the molecular bonds
46: is obtained.
47: 
48: With spectroscopy one can study properties of materials and
49: mixtures without destructing the sample. There are crudely speaking
50: two kinds of spectra, relying on
51: different experimental methods. An emission spectrum is obtained by
52: putting a system in a state of high energy. The system then falls
53: back to a state with lower energy, and the energy difference is
54: emitted in the form of light. Of course, in order to emit light, the
55: system needs to interact with light. The kind of
56: interaction then dictates which transitions are possible and hence
57: which frequencies are emitted. For the absorption spectrum one more or
58: less does the converse. One puts a system into a beam of (nearly) white
59: light. The system then absorbs light and re-emits it again, but then
60: in all directions.
61: 
62: In the 19th century Kirchhoff used an invention of the German chemist 
63: Robert Bunsen to heat up elements in a flame to study the emitted
64: light. He passed light through a prism to study 
65: the intensity of light at different wavelengths. It turned out
66: that the emitted spectrum of an element had
67: quite clearly defined lines at certain wavelengths. In 1859
68: Kirchhoff pointed out that all the elements that he had been studying
69: had a different emission spectrum. Hence
70: disentangling the lines of an emission spectrum can help in finding the
71: components an unknown mixture is made of. Figure \ref{he-spec} gives 
72: as example an emission and an absorption spectrum of Helium.
73: 
74: \begin{figure}[htb]
75: \begin{pspicture}(1,-0.5)(18,4.5)
76: \psspectrum[element=He](1,2.2)(18,4.2)
77: \psspectrum[absorption,element=He](1,0)(18,2)
78: \end{pspicture}
79: \caption{The emission (upper) and absorption (lower) spectrum of
80: Helium.}\label{he-spec}
81: \end{figure}
82: 
83: 
84: Already much earlier, Isaac Newton had used in 1670-1672 a prism to 
85: study the decomposition of
86: white light into a spectrum of different colors. In 1814 Joseph von
87: Fraunhofer invented the spectroscope and identified 574 dark lines in
88: the light of the sun. In fact, the Fraunhofer experiment can already
89: be done with primitive equipment. On a sunny, cloudless day one
90: sits in a dark room with one little hole through which the sun
91: shines. In the beam of sunlight one places a prism and lets the
92: light after the prism fall onto a white piece of paper. The observed
93: spectrum can be seen to display dark lines; in Figure \ref{he-spec}
94: the lines corresponding to helium are displayed. The Fraunhofer lines
95: are a manifestation of the absorption 
96: spectrum. It was Kirchhoff who later explained the origin; light from
97: the sun has to pass the atmosphere of the sun. In the atmosphere the
98: elements that are present absorb certain parts of sunlight, at
99: well-defined frequencies and re-emit it later, but then in all
100: directions. Therefore the sunlight going in the forward direction --
101: that is, away from the core of the sun -- has lost intensity at certain
102: well-defined frequencies. In this way, Kirchhoff showed that the
103: atmosphere of the sun contained among others hydrogen and sodium.
104: The reason why the sunlight is almost white before
105: entering the atmosphere of the sun we will not explain. When the light
106: of the sun reaches the earth it is already so diluted that the
107: elements in the earths' atmosphere give almost unobservable absorption
108: lines. Therefore, the dark lines in the spectrum of the sun are due to
109: the suns' atmosphere and not the earths' atmosphere.
110: 
111: In 1868, the French astronomer Pierre-Jules-Cesar Janssen observed a
112: line in the spectrum of the sun that did not match any element known
113: by then.
114: The reason he observed it and not Fraunhofer was because Janssen
115: used the better observing circumstances that a solar eclipse
116: offers. Normally, the sun is too bright, but when the moon blocks the
117: solar disc, one sees solely the atmosphere of the sun. The astronomer
118: Joseph Norman Lockyear concluded that the new line must represent a
119: new element. They tossed the name helium, from the Greek word
120: ``helios'', which means sun. It was not until 1895 that the physicist
121: John William Strutt, Lord Rayleigh -- or in short John Rayleigh --
122: proved that helium is also present on earth; he found it in samples of
123: the mineral clevite. He exposed the mineral to some
124: acids that reacted with the material thereby producing gasses. Then
125: he studied the contents of the gas mixtures, and he found that helium
126: was present. The reason why he found helium was
127: explained later. Clevite is a mineral that contains uranium. The element
128: uranium is radio-active; it can emit $\alpha$-particles, which are the 
129: nuclei of helium atoms. 
130: 
131: 
132: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
133: \section{Spectra of systems of particles}
134: 
135: We distinguish two kinds of spectra:\\
136: 1. The spectrum in the sense of spectroscopy is the collection of
137: frequencies emitted or absorbed by the system in its interaction with
138: light or other electromagnetic (infrared, radio, X-ray) radiation.\\
139: 2. The spectrum of a physical system is the collection of allowed
140: energy values -- the set of eigenvalues of the associated Hamiltonian.
141: 
142: The relation between the two is as follows. The observed
143: spectrum (of spectroscopy) consists of the energy differences of the
144: system: the observed spectra are of the form $\hbar
145: \omega_{mn}=E_m-E_n$, where the energy levels of the system are
146: $E_n$. In most systems the spectrum is discrete. Hence also the
147: observed spectrum is discrete.
148: 
149: For systems that
150: are made of  constituents that can break apart, the spectrum contains
151: continuous parts. Consider for
152: example a molecule of two atoms like $H_2$. At a
153: certain frequency the molecule can break apart. Then the energy of
154: the photon can also be put into the kinetic energy of both $H$-atoms,
155: which is a continuous parameter. 
156: 
157: 
158: If the Hamiltonian has some imaginary eigenvalues $\lambda$, then $\im
159: \lambda<0$ and the modes corresponding to $\lambda$ are decaying
160: modes. In a dissipative environment this results in energy loss, and
161: the system can move from higher energy to lower energy.
162: 
163: On the other hand, a system can also be excited. It then absorbs
164: energy from the environment. A typical example of
165: excitation is an atom interacting with light. The energy levels of the
166: atom are discrete, and hence only with a fine-tuned frequency the atom
167: can absorb a photon and attain a state with more energy. The energy
168: difference between the ground state and the state with the second
169: lowest energy is called the energy gap. If a photon has the frequency
170: with the energy corresponding to the energy gap, it can be absorbed by
171: the atom and the atom can be excited to the state above the ground
172: state.
173: 
174: 
175: An excited atom cannot move down to a state with lower energy due to
176: energy conservation, unless there is no interaction with light.
177: Incorporating interaction with light into the Hamiltonian makes the 
178: energies acquire a
179: small imaginary part, representing the possibility to decay. If an
180: atom jumps down in energy, it emits a photon with the same
181: energy. This process is called spontaneous emission. The nice feature
182: of spontaneous emission is that we can observe it.
183: 
184: The interaction with light is not just any arbitrary interaction. The
185: interaction term $V_\fns{int}$ in the Hamiltonian
186: \[
187: H_\fns{tot}=H_\fns{atom}+H_\fns{env}+V_\fns{int}
188: \]
189: needs to respect some symmetries
190: like Galilean invariance. \at{explain or avoid}
191: The result is that not all transitions
192: but only a selected set of transitions is allowed. The rules that
193: dictate which transitions are allowed are therefore called
194: {\bfi{selection rules}}.
195: 
196: The interaction is often treated as a perturbation. The justification
197: is that the interaction term in the Hamiltonian is small compared to
198: the other terms. One introduces a
199: dimensionless variable $\lambda$ and re-writes
200: $V_{int}$ as $V_{int}(\lambda)= \lambda V_{int}$. One recalculates the
201: spectrum and expands it in $\lambda$ to find
202: \[
203: E_k(\lambda)= E_k(0)+\lambda \Delta E_{k}^{1}+\lambda^2 \Delta
204: E_{k}^{2}+\ldots\,.
205: \]
206: Since the interaction is small, the first order correction often
207: gives the interaction with light accurately enough.
208: Using the techniques of perturbation theory
209: one then finds the possible transitions, i.e. the selection rules, and
210: the probabilities of the transitions. The probabilities gives the
211: dominance in the observed spectrum; if a transition A is more probable
212: than a transition B this will result in more spontaneous emission
213: along transition A. Therefore the peak in the spectrum corresponding
214: to A is bigger than the peak corresponding to B.
215: 
216: 
217: 
218: \begin{figure}
219: \begin{center}
220: \includegraphics[scale=0.35]{spectrum-ancker2.eps}
221: \caption{An example of a spectrum.}\label{spec1}
222: \end{center}
223: \end{figure}
224: 
225: 
226: Observed spectra are often displayed by plotting, 
227: as in Figure \ref{spec1}, \at{poor figure quality}
228: on the horizontal
229: axis the frequency and on the vertical axis the observed
230: intensity. Due to imperfections in measuring methods one never
231: observes a real peak, but always a smeared out peak, that is, peaks
232: have a width. However, there can be many reasons why a peak has a
233: certain width. Imagine for example that one measures the spontaneous
234: emission of a gas contained in  cylinder. The gas atoms are moving
235: around in the cylinder, with different velocities with respect to the 
236: measuring device. For each atom the spectrum is shifted
237: due to the Doppler effect, known from a similar effect with sound,
238: which can be observed when an ambulance passes by. Since one measures 
239: the emission of a whole population of atoms, the measured peak is a 
240: superposition of peaks that are distributed around a certain frequency. 
241: That is, the Doppler effect broadens a peak. 
242: 
243: Technical imperfections of the measuring device also broaden peaks.
244: Making the measuring equipment more and more accurate one can try to 
245: get a better and better resolved spectrum. Doing this one might see that
246: broad peaks resolve into a group of smaller peaks. One sees therefore
247: more structure.
248: 
249: The result of a measurement is a list of data, the
250: frequencies $\omega_l$. Using the data one wants to obtain information
251: of the system under study. If one knows the system already quite well,
252: for example if one knows the parametric form of the Hamiltonian but 
253: not the precise values of the
254: parameters, one may fit the measured energies to obtain a set of
255: parameters that describes the measurements best. One therefore 
256: has to solve a data analysis problem.
257: For each label $l$ one has to find energies $E_{k(l)}$ and $E_{j(l)}$
258: with
259: \[
260: \omega_l \approx E_{j(l)} - E_{k(l)},
261: \]
262: within the experimental accuracy. Therefore, one solves the
263: least-squares problem of minimizing the sum
264: \[
265: S(E,j,k):=
266: \sum_l q_l \Big( \frac{E_{j(l)} - E_{k(l)}}{\hbar \omega_l} -1\Big)^2\,,
267: \]
268: for some weight factors $q_l $ related to the inverse of the accuracy
269: of the measurement of $\omega_l$.
270: 
271: In general, both the list $E$ of energy levels $E_i$ and the functions
272: $j,k$ which determine the {\bfi{assignment}} of spectroscopic 
273: lines to
274: transitions are unknown, and must be determined by minimizing
275: $S(E,j,k)$. Usually, one starts with a preliminary list $E$ of energy
276: levels, and assigns each line $l$ to a transition which minimizes
277: the $l$th term, breaking ties arbitrarily. This defines preliminary
278: assignment functions $j,k$. Fixing these turns the problem of minimizing
279: $S(E,j,k)$ into a least squares problem for finding the energy levels,
280: resulting in an improved $E$. Clearly, each cycle decreases the
281: value of $S(E,j,k)$. The process is stopped when the assignments
282: no longer change. Then $S(E,j,k)$ has reached a local minimum.
283: Multiple lists of trial energy levels may be used to increase the
284: likelihood that the assignment found corresponds to a global minimum.
285: Frequently, one first assigns a subset of lines to a subset of levels
286: to find good starting values.
287: 
288: 
289: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
290: \section{Examples of spectra}
291: 
292: The geometry of the molecule or atom under consideration strongly
293: influences the spectrum, since the geometry determines the potential.
294: 
295: \begin{figure}
296: \begin{center}
297: \includegraphics[scale=0.7]{harm-osc3.eps}
298: \caption{Harmonic oscillator potential, with the eigenvalues indicated.}
299: \label{spec2}
300: \end{center}
301: \end{figure}
302: 
303: Consider a molecule of two atoms. We assume that the excitations
304: inside each atom are of another magnitude than the excitations of the
305: bond between the atoms. In that case we may consider the molecule as
306: two balls connected by a spring. The spectrum is as in
307: Figure \ref{spec2}, and the observed spectrum consists of one peak.
308:  \at{poor figure quality}
309: 
310: 
311: Consider now a system that has two local minima. An example of this
312: would be a molecule $C_2H_4$ of which two versions exist, the cis and
313: trans molecules. The molecular bond between the two $C$-atoms then
314: behaves around each local minimum as a harmonic oscillator in some
315: approximation. For higher energies however the two states start to
316: interact and the molecule can change from cis to trans and vice
317: versa. A typical spectrum then looks like Figure \ref{spec3}.
318: 
319: \begin{figure}
320: \includegraphics[width=0.4\linewidth]{doublwell3.eps}
321: \hfill
322: \includegraphics[width=0.50\linewidth]{dissociation2.eps}
323: \caption{On the left a double well potential with the first energy 
324: levels indicated. On the right the Morse potential; 
325: the bound states have discrete energy but above the dissociation 
326: energy the spectrum is continuous.}\label{spec3}
327: %\end{center}
328: \end{figure}
329: 
330: 
331: When there are asymptotically free states, one says that the system
332: admits {\bfi{dissociation}}. Free states have continuous kinetic
333: energy and hence the spectrum contains continuous parts. 
334: A potential showing
335: dissociation is the Morse potential given by
336: \[
337: V(r) = \alpha(e^{-\beta r} - \gamma)^2 -\alpha\gamma^2\,, r\geq 0
338: \]
339: where $r$ is the atomic distance and $\alpha$, $\beta$, and $\gamma$ 
340: are positive parameters, see Figure \ref{spec3}. 
341: The potential of the $H_2$ molecule discussed above is another example.
342: Above the dissociation energy the spectrum is continuous; the bound 
343: states have discrete energy.
344: 
345: %\at{in Figure \ref{spec4}, change $E$ to $V(r)$.}
346: %\begin{figure}
347: %\begin{center}
348: %\includegraphics[scale=0.7]{dissociation2.eps}
349: %\caption{The Morse potential. The bound states have discrete energy. 
350: %Above the dissociation energy the spectrum is continuous.}\label{spec4}
351: %\end{center}
352: %\end{figure}
353: 
354: Quantum physics has a remarkable feature compared to classical
355: mechanics, called {\bfi{tunneling}}. 
356: If a particle is in a local minimum at
357: energy $E_1$ and another minimum is available with energy level
358: $E_0 < E_1$, then there is a nonzero probability that the particle
359: `travels through the barrier' and ends up in the local minimum with
360: lower energy. For example, the potential of the $C_2H_4$-molecule 
361: discussed above admits tunneling since the potential has two local
362: minima. The probability of tunneling decreases with the height of the
363: barrier between the two energy levels. Another example where tunneling
364: occurs is in nuclear physics; the potential of Figure \ref{spec5}
365: represents the energy a proton feels in the potential field of a 
366: nucleus. The diameter of the nucleus is roughly the distance between 
367: the two peaks in Figure \ref{spec5}. The difference to the 
368: $C_2H_4$-molecule is here that the tunneling takes place between two
369: states one of which is not integrable. Tunneling can go in two
370: different directions; one direction is where the proton is shot 
371: at the nucleus with too little energy to classically penetrate the
372: nucleus, the other direction is where the proton is inside the nucleus
373: and classically cannot get out. In the latter case, there is a certain
374: probability that the proton escapes the nucleus. This explains
375: qualitatively the stochastic behavior of radio-active decay. 
376: 
377: \begin{center}
378:  \begin{figure}
379:   \input{barrierpot.pstex_t}
380:   \caption{Sketch of the potential a proton experiences in the force
381:  field of a nucleus.}\label{spec5}
382:  \end{figure}
383: \end{center}
384: 
385: \at{the figure produces a spurious space in the text!}
386: As another example, consider a chemical reaction of the form
387: $AB+C\to A+BC$, that is, the molecule $AB$ splits off a part $B$ that
388: then attaches to $C$ to form $BC$. Here there are two important
389: parameters. The distance $|AB|$ between $A$ and $B$ and the
390: distance $|BC|$ between $B$ and $C$. A possible potential is plotted
391: in Figure \ref{spec6}. The plot shows two valleys separated 
392: by a saddle point, marked by a red cross. 
393: \at{red not visible in print; also in the caption}
394: The horizontal valley 
395: corresponds to $|BC|$ constant, hence to the state $A+BC$. 
396: The other valley corresponds to $AB+C$, and at the saddle point part 
397: $B$ is exchanged.
398: 
399: 
400: \begin{figure}
401: \begin{center}
402: \includegraphics[scale=0.5]{heightlines2.eps}
403: \caption{ 2D-Plot of the potential experienced in a chemical reaction 
404: $AB+C\to A+BC$. The potential 
405: depends on the distances $|AB|$ and $|BC|$. The upper part shows a 
406: cross-section of the potential. The red marker is the saddle 
407: point.}\label{spec6}
408: \end{center}
409: \end{figure}
410: 
411: \at{Figure \ref{spec6} is black in print!} 
412: 
413: 
414: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
415: \section{Dynamical symmetries}\label{s.dyn}
416: 
417: As discussed before, when one looks at a poorly resolved spectrum,
418: one sees some rough features of the system under study. Improving the
419: resolution allows one to study more structure of the system.
420: 
421: A similar process happens when one studies a hydrogen atom in an
422: external magnetic field. Upon increassing the magnetic field one sees 
423: that many lines of the original spectrum split into several close
424: lines. Thus what first seems to be one state in
425: fact turns out to be an agglomeration of different states. The states
426: first had energies that were so close together that they could not be
427: recognized as belonging to different states -- indeed, they have
428: exactly the same energy. As we shall see, that these states (seemingly)
429: agglomerate to one single state is due to symmetry reasons.
430: 
431: The rotational symmetry implies that the energies of different states
432: related by a rotation have the same energy; more pictorially, whether
433: an electron circles around the proton with the rotation axis in the
434: $z$-direction or the $y$-direction gives the same energy.
435: Turning on the magnetic field results in breaking the symmetry;
436: then the different states that first agglomerated to form a single
437: state are disentangled and can be observed separately in the spectrum.
438: 
439: But as with the increasing resolution, taking a closer look at the
440: hydrogen atom reveals more and more structure. In a first
441: approximation, the electron in the hydrogen atom can be treated
442: nonrelativistically. Treating the electron relativistically, one gets
443: a correction to the spectrum. The first order corrections of special
444: relativity go under the name of the {\bfi{first radiative 
445: corrections}}.
446: 
447: We shall look in some detail at the hydrogen atom once we have 
448: clarified the general principles.
449: 
450: \bigskip
451: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
452: \bfi{Symmetry and broken symmetry.}
453: The most symmetric physical systems, in particular the standard 2-body
454: problems (the classical Kepler problem and the quantum hydrogen atom)
455: are exactly solvable. The Helium atom is already a three-body problem
456: and is not exactly solvable.
457: 
458: A physical system is called {\bfi{exactly solvable}} 
459: (or {\bfi{integrable}}, or {\bfi{completely  integrable}}) 
460: if it has ``enough'' constants of motion. Equivalently, if the 
461: centralizer $C_\Ez(H)$ of the Hamiltonian $H$ in the algebra $\Ez$ 
462: of observables is ``large enough''. 
463: The effect of having enough central elements is that the
464: system has enough conserved quantities to explicitly solve the
465: differential equations of the system. 
466: \at{give references to exact definitions in the classical case}
467: 
468: A {\bfi{dynamical algebra}} of a classical physical system is a Lie
469: algebra $\Lz$ that one can associate to the system such
470: that the Hamiltonian $H$ is contained in the Lie--Poisson
471: algebra $ C^\infty(\Lz^*)$. For a quantum mechanical system the
472: corresponding requirement is that $H$ is contained in the closure of 
473: the universal enveloping algebra $\mathcal{U}(\Lz)$ of $\Lz$, equipped 
474: with a locally convex topology such that
475: potentials of the form $e^{-x^2}$ are allowed.
476: 
477: For example, the Heisenberg algebra $h(n)$ is the dynamical
478: algebra of symplectic classical systems with $n$ position degrees of
479: freedom, and of traditional Schr\"odinger quantum mechanics.
480: The hydrogen atom has additional rotational symmetry, and the
481: special properties of the Coulomb potential imply that one can in
482: fact find a fairly big dynamical algebra, namely $so(2,4)$, see
483: e.g. \sca{Wybourne} \cite{wybourne}. 
484: 
485: 
486: Now consider any Lie algebra $\Lz$ as a dynamical algebra. Call $\Ez$
487: the Lie--Poisson algebra associated to $\Lz$ for a classical case or
488: the universal enveloping algebra of $\Lz$ in the quantum case. The {\bf
489: \idx{symmetry algebra}} is the centralizer of the Hamiltonian in $\Ez$,
490: written $C_\Ez(H)$. In the `nicest' case one has $\Ez=C_\Ez(H)$, which
491: means that $H$ is a Casimir of $\Lz$. Normally, the Lie algebra
492: $\Lz$ describes the symmetries of the (unperturbed) system and thus
493: one would expect that the nicest case is the general case.
494: 
495: However, a very symmetric system is rarely studied in isolation, and
496: realistic systems are at best perturbations of nice systems. In this
497: case one gets broken symmetries, meaning that the Hamiltonian is only
498: almost a Casimir. Note that it might happen that the classical theory
499: has a symmetry, but that in the quantum version of the theory the
500: symmetry gets broken.
501: In case of a broken symmetry, one usually first tries to solve the
502: symmetric problem and then perturb the solutions to get approximate
503: solutions to the problem with broken symmetry.
504: We will not go into details about the mathematics of perturbation
505: theory, since this topic is amply treated in every book on quantum
506: mechanics. But we will consider some of its qualitative implications.
507: 
508: Suppose we have solved a symmetric problem. Then the solutions are
509: described as elements of some Hilbert space $\Hz$ on which $\Lz$ acts
510: unitarily. We can decompose the
511: Hilbert space into a direct sum of eigenspaces of the Hamiltonian;
512: $\Hz=\oplus_\lambda \Hz_\lambda$. Let $\psi$ be some eigenstate in
513: $\Hz_\lambda$ of the Hamiltonian and let $f\in \Lz$, then we see
514: \[
515: Hf\psi = [H,f]\psi + fH\psi =  [H,f]\psi +\lambda f\psi\,.
516: \]
517: Since $[f,H]=0$ we see that $\Lz$ maps each eigenspace into
518: itself. Thus all $\Hz_\lambda$ are $\Lz$-modules.
519: 
520: We call the eigenvalue $\lambda$ {\bf
521:   nondegenerate}\index{nondegenerate eigenvalue} if the dimension 
522: of $\Hz_\lambda$ is 1, and \bfi{degenerate}\index{degenerate eigenvalue}
523: if it is bigger than 1. (Dimension zero means that $\lambda$ is not an
524: eigenvalue.) 
525: 
526: 
527: If $\lambda$ is degenerate, $\Hz_\lambda$ has
528: many essentially distinct bases of eigenvectors of $H$. One of
529: these is usually distinguished by the concrete representation used to
530: describe the module; $H$ and usually a distinguished
531: part of $\Lz$ act diagonally. In general the perturbed Hamiltonian no
532: longer acts 
533: diagonally on $\Hz_\lambda$, and as a result the level $\lambda$
534: usually splits into several distinct levels.
535: The energy level $\lambda$ splits into new levels that
536: are of the form $\lambda+\epsilon_j$ for some different but small
537: values $\epsilon_j$, giving rise to a fine structure. In
538: general, a fine structure implies that either a symmetry is broken
539: (the system reached a nonsymmetric state) or an external force that
540: broke the symmetry explicitly has been applied.
541: 
542: The induced representation of $\Lz$ on $V_\lambda$ is
543: unitary. Therefore knowing the irreducible unitary representations of
544: $\Lz$ can give information about the system under study.
545: 
546: 
547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
548: \section{The hydrogen atom}
549: 
550: A hydrogen atom is a bound state of a proton (the nucleus) and an 
551: electron. It is most esaily described by treating the much heavier 
552: nucleus as fixed (which amounts to neglecting recoil effects) and 
553: considering the electron as moving in the electrostatic Coulomb field 
554: generated by the nucleus.
555: 
556: The electron is a spin $1/2$ particle, a fermion, meaning that it is
557: described by the spin $1/2$ representation of $so(3)$ on the
558: Hilbert space $L^2(\Rz^3,\Pz_1) \cong L^2(\Rz^3)\otimes\Pz_1)$ 
559: defined in Section \ref{s.spinor}. 
560: Below we first discuss the orbital part of the wave
561: functions, i.e. the $L^2(\Rz^3)$-part. Then we discuss the dynamical
562: symmetries and how they get broken.
563: 
564: The orbital quantum states are
565: labeled by integers $n$, $l$ and $m$. The integer $n$ takes the
566: values $1,2,3,\ldots$ and the number $l$ takes for each fixed
567: value of $n$ the values $0,1,2,\ldots,n-1$. Finally, the number $m$
568: takes for each $l$ the values $-l,-l+1,\ldots,l-1,l$. 
569: Hence the (orbital) state of an electron is described by a state
570: \begin{equation}
571: \vert n,m,l\rangle ~~~ {\rm where}~~~ n\geq 1\,, ~~
572: 0\leq l< n\,,~~ -l\leq m\leq l\,.
573: \end{equation}
574: The {\bfi{quantum number}} $n$ determines (to a first approximation)
575: the energy of the state:
576: \begin{equation}
577: E_n = -\frac{13.6 eV}{n^2}\,.
578: \end{equation}
579: The abbreviation $eV$ means electron Volt and is a unit for
580: energy. The quantum number $l$ specifies a representation of
581: $so(3)$. Thus we can make use of the representation theory of
582: $so(3)$ developed in Section \ref{s.su2rep}.
583: 
584: The electrostatic potential of the
585: hydrogen atom is $SO(3)$-invariant, hence it is not too surprising
586: that $SO(3)$-representations plays a role; the orbital part of the
587: electron wave function can be decomposed in representations of $SO(3)$.
588: The quantum 
589: number $l$ corresponds precisely to the irreducible representation
590: $D_l$ of $so(3)$, that is, precisely to the representations of $so(3)$
591: that lift to $SO(3)$-representations. The quantum number $m$ labels 
592: the $\sigma_3$-eigenvectors of the representation and
593: corresponds to the eigenvalue $m$. 
594: The quantum number $n$ thus determines
595: which $SO(3)$-representations are allowed, and the $l$ and $m$ then
596: specify the representation and an eigenvector in this representation.
597: 
598: Now we shortly describe the relation between the quantum numbers and
599: the orbital wave function of the electron in the hydrogen atom. We can
600: give the hydrogen atom a coordinate system as follows. We put the
601: proton in the center and describe the position of the electron by a
602: radial coordinate $r$ measuring the distance between the proton and the
603: electron and by two angles $\theta\in [0,\pi ]$ and $\phi\in
604: [0,2\pi)$. The solutions to the Schr\"odinger equation for the
605: hydrogen atom are then given by
606: \[
607: \psi(r,\theta,\phi) = R_{n,l}(r)Y_{l,m}(\theta,\phi)\,.
608: \]
609: The radial part of the wave function $R_{n,l}$ is completely
610: determined by the quantum numbers $n$ and $l$ and is given by
611: \[
612: R_{n,l}=C_{n,l}e^{-\rho/2}\rho^l L_{n-l-1}^{2l+1}(\rho).
613: \]
614: Here $C_{n,l}$ is some constant such that $R_{n,l}$ is
615: normalized to integrate to one, the $L_{q}^{p}(\rho)$ are
616: generalized Laguerre polynomials (one of the well-known families of 
617: special functions);
618: $\rho $ is the normalized radius $\rho = \frac{2r}{na_0}$, and
619: $a_0$ is a constant called the {\bfi{Bohr radius}}. 
620: The angular part $Y_{l,m}$ of the wave function is given by
621: \[
622: Y_{l,m}(\theta,\phi) = K_{l,m}P_{l}^{m}(\cos \theta)e^{im\phi}\,,
623: \]
624: where the $K_{l,m}$ are normalization constants, and the $P_{l}^{m}$ 
625: are the {\bfi{associated Legendre polynomials}} given by
626: \[
627: P_{l}^{m}(x)=\frac{(-1)^m}{2^ll!}(1-x^2)^{m/2}
628: \left(\frac{d}{dx}\right)^{l+m}(x^2- 1)^l\,.
629: \]
630: 
631: \bigskip
632: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
633: \bfi{Symmetries and symmetry breaking}.
634: Nonrelativistically the electron in an electromagnetic field is treated
635: with the {\bfi{Pauli equation}}. The Pauli equation looks like the
636: Schr\"odinger equation, but has some extra terms, describing the
637: coupling of a spin $1/2$ particle to the electromagnetic
638: field. We now indicate why, in the case where the external
639: electromagnetic field is switched off, the symmetry group of the
640: Hamiltonian is $SO(4)\times SO(3)$.
641: 
642: The second factor in the symmetry group, the $SO(3)$, is the
643: symmetry group that acts on the spin of the electron. That is, it acts
644: on the $D_{1/2}$ part of $L^2(\Rz^3)\otimes D_{1/2}$.  
645: 
646: 
647: The first factor in the symmetry group, the $SO(4)$, acts on the
648: space-part $L^2(\Rz^3)$ of the wave function. The Hamiltonian
649: of the hydrogen atom is rotationally invariant. 
650: Infinitesimal rotations are generated by the angular momentum 
651: $\L = \mathbf{r}\times\mathbf{p}$, where $\mathbf{r}$ is the
652: radius and $\mathbf{p}$ is the linear momentum and hence the angular 
653: momentum
654: components describe the Lie algebra $so(3)$. However, there exists an
655: additional vector whose length is conserved: the
656:  length of the {\bfi{Lenz--Runge vector}}. (Some people call it the
657: Laplace--Runge--Lenz vector, or even Laplace vector.)
658: This leads to the bigger group $SO(4)$;  see e.g. \sca{Goldstein}
659: \cite{goldstein}. 
660: 
661: 
662: To treat the electron relativistically one uses the Dirac
663: equation for a spin $1/2$ particle coupled to an electromagnetic
664: field. The coupling to the electromagnetic field can be done
665: in a quite easy way. Starting with the Dirac equation
666: \[
667: \left( \hbar\gamma\cdot \partial + mc\right)\psi=0
668: \]
669: one simply replaces the derivatives with $\partial_\mu - iqA_\mu$
670:  where the zeroth component of $A$ gives the Coulomb
671: potential and the spatial $\mathbf{A}$ components contain the magnetic
672: field via $B = \Nabla\times \mathbf{A}$; the parameter $q$ is
673: interpreted as the charge. We obtain
674: \[
675: \left( \hbar\gamma\cdot \partial -iq\hbar \gamma\cdot A +
676:   mc\right)\psi=0\,,
677: \]
678: where $\gamma\cdot A=\gamma^\mu A_\mu$.
679: 
680: The effect of having the fully relativistic coupling terms is that 
681: there is a coupling between the spin of the electron and the
682: orbital angular momentum of the electron.
683: The additional coupling terms in the
684: Hamiltonian are called {\bfi{spin-orbit coupling} terms}. Due to the
685: coupling the separate $SO(3)$ of the spin gets destroyed; without
686: coupling there is a rotational symmetry group acting separately on the
687: orbit and on the spin and due to the coupling, the two rotational
688: symmetries are no longer independent. The angular momentum
689: $\mathbf{L}$ and the spin $\mathbf{S}$ are no longer separately
690: conserved in magnitude, but $(\mathbf{L}+\mathbf{S})^2$ is
691: constant. The symmetry group of the
692: relativistic hydrogen atom is therefore $SO(4)$. The spectrum that is
693: observed is called the {\bfi{fine structure spectrum}}.
694: 
695: 
696: Going even further and treating the hydrogen atom with quantum field
697: theory results in a further breakdown of the symmetry to the group
698: $SO(3)$. The group $SO(4)$ is isomorphic\footnote{This isomorphism is
699:   proved along the same lines displayed in Section \ref{s.rot};
700:   first one proves $so(4)\cong so(3)\oplus so(3)$ and then notes that
701:   $so(4)$ is connected.} to $SO(3)\times SO(3)$ and
702: corrections from quantum field theory break it down to the diagonal
703: subgroup $SO(3)$. The observed spectrum is called the {\bf
704:   \idx{hyperfine structure spectrum}}. 
705: 
706: 
707: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
708: \section{Chains of subalgebras}\label{s.chains}
709: 
710: In more realistic situations,  the Hamiltonian is not
711: invariant under the total dynamical algebra $\Ez$, the universal
712: enveloping algebra of $\Lz$. In this case, the
713: Hamiltonian is not in the center of $\Ez$, but we can consider the
714: centralizer of $H$ in $\Lz$. The centralizer of $H$ in $\Lz$ is a
715: subalgebra of $\Lz$, and is therefore a Lie subalgebra of $\Ez$ and we
716: denote it by $\Lz_1$. We thus have $H\in C_{\Ez}(\Lz_1)$. The Lie
717: subalgebra $\Lz_1$ generates a subalgebra of $\Ez$, which we denote by
718: $\Ez_1$.
719: 
720: In simple applications, it often happens that the Hamiltonian $H$ is a
721: function $H(C_0,C_1)$ where $C_0$ is a Casimir of $\Lz$ (that is, it
722: is a central element of $\Ez$) and where $C_1$ is a Casimir of $\Lz_1$
723: (in the center of $\Ez_1$). In more complicated applications we have a
724: series of approximations to the problem, as explained for the hydrogen
725: atom before, where relativistic and quantum field theory effects
726: modify the Hamiltonian. In each step one modifies the
727: Hamiltonian by adding terms with ferer and ferwer symmetries, and the 
728: symmetry algebra is reduced to correspondingly smaller
729: subalgebras. We thus have a sequence of subalgebras
730: \[
731: \Lz = \Lz_0 \supseteq \Lz_1\supseteq \ldots \supseteq \Lz_n=\hat
732: \Lz\,.
733: \]
734: The final subalgebra $\hat \Lz$ commutes with $H$. 
735: The generated subalgebra
736: of $\Ez$, denoted $\hat \Ez$ centralizes $H$ in $\Ez$. If the
737: Hamiltonian is a function $H=H(C_0,\ldots,C_n)$ where $C_k$ is a
738: Casimir of $\Lz_k$, the scheme gives
739: explicitly solvable problems. For example, for the nonrelativistic
740: hydrogen atom without spin, one finds a series
741: \[
742: so(4)\supset so(3) \supset so(2)\supset 1 \,.
743: \]
744: Of course, there are many Hamiltonians that cannot be represented
745: as functions of a chain of Casimirs, but the above scheme covers
746: many applications, and is a starting point for a perturbative treatment
747: of many others.
748: 
749: In classical symplectic mechanics one relates the Lie algebra
750: $\hat \Lz$ to so-called {\bfi{action variables}} and the steps to
751: $\Lz_0$ are constructed using conjugate {\bfi{angle variables}}.
752: We will not go into the details defining variables and the related 
753: techniques. \at{refs}
754: 
755: 
756: Consider the situation where $H=H(C_0,C_1)$, that is, the
757: simple application. We write $H=H_0+H_1$ where $H_0$ is only a
758: function of $C_0$ and $H_1$ depends on $C_0$ and $C_1$. As before,
759: we suppose we have realized the
760: elements of $\Ez$ (and thus of $\Lz$) as operators on some Hilbert
761: space $\Hz$. We assume that the subspaces $\Hz_\lambda$ on which the
762: Hamiltonian $H_0=H_0(C_0)$ acts diagonally are finite-dimensional.
763: This is for example the case for the hydrogen atom. We furthermore
764: split up $\Hz_\lambda$ in irreducible representations of $\Lz_0$
765: so that we may assume that $\Hz_\lambda$ is irreducible.
766: Modifying $H_0$ to $H_0+H_1$ means that the symmetry algebra becomes
767: smaller; it 
768: becomes $\Lz_1$. We can restrict the representation of $\Lz_0$ on
769: $\Hz_\lambda$ to the subalgebra $\Lz_1$ to obtain a representation of
770: $\Lz_1$. In most cases this representation is reducible and we
771: write the decomposition of $\Hz_\lambda$
772: into $\Lz_1$ irreducibles as
773: \[
774: \Hz_\lambda  = \bigoplus_{\mu} \Hz_{\mu}^{(1)}\,.
775: \]
776: More generally, suppose we have a sequence of subalgebras $\Lz_0\supset
777: \Lz_1\supset \ldots\supset \Lz_n$ and related Casimirs $C_i\in
778: Z(U(\Lz_i))$. It follows that the $C_i$ commute among each other;
779: $C_1$ is in the center of $U(\Lz_1)$, which contains $U(\Lz_k)$ for
780: $k\geq 2$ and hence $C_1$ commutes with $C_2$, $C_3$, and so on. Thus
781: on the irreducible representations of $\Lz_1$ the Casimirs act
782: diagonally. Hence we can assign to each representation of $\Lz_n$
783: appearing in the decomposition of the original representation $\Hz$
784: numerical values $\theta_0, \ldots, \theta_n$, corresponding to the
785: eigenvalues of the $C_i$. Given a physical state $v$ in a
786: representation of $\Lz_n$ corresponding to the label
787: $(\theta_0,\ldots,\theta_n)$ we see that the Hamiltonian acts as
788: \[
789: H(C_0,\ldots,C_n)v= H(\theta_0,\ldots,\theta_n) v\,,
790: \]
791: where on the right-hand side the Hamiltonian is an operator and on the
792: right hand side $ H(\theta_0,\ldots,\theta_n)$ is a number.
793: 
794: 
795: \bigskip
796: %%%%%%%%%%%%%%%%%%%%%%%
797: {\bfi{Branching rules}.}
798: In a lot of favorable cases, for example when the $\Lz_i$ are simple,
799: the decomposition of an irreducible representation of the large
800: algebra into irreducible representations of a maximal subalgebra
801: is known. These decompositions go under the name of \bfi{branching
802: rules}; splitting up a representation
803: under a subgroup or subalgebra is called \bfi{branching}.
804: 
805: Let us give an example of a branching rule and look at the
806: fundamental representation of $su(3)$, that is, the $3$-dimensional
807: representation of $su(3)$ that defines the Lie algebra $su(3)$. 
808: The Lie algebra elements are faithfully represented as $3\times
809: 3$-matrices $X$ that are antihermitian; $X^\dagger+X=0$. Now we
810: consider the Lie subalgebra $su(2)$. There are different ways we can
811: embed $su(2)$ into $su(3)$, but it turns out that all of them are
812: equivalent. We can always choose a basis $e_1$, $e_2$, $e_3$ in
813: $\Cz^3$ such that $su(2)$
814: only acts nontrivially on the subspace spanned by $e_1$ and $e_2$. We
815: thus realize $su(2)$ inside $su(3)$ as the following matrices in
816: $su(3)$
817: \[
818: \pmatrix{ a & b & 0 \cr c & d& 0 \cr 0 & 0 & 1  }\,,~~~\mbox{where}~~
819: \pmatrix{ a & b \cr  c & d  }\in  su(2)\,.
820: \]
821: We see that the three-dimensional representation
822: of $su(3)$ splits into two irreducible representations of $su(2)$, the
823: trivial one, spanned by $e_3$ and the two-dimensional (fundamental)
824: representation spanned $e_1$ and $e_2$. One writes this in
825: shorthand as: $\mathbf{3\to 2+1}$ under $su(2)\subset su(3)$. In the
826: reference \sca{Slansky} \cite{slansky}, one can find tables of
827: branching rules. 
828: 
829: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
830: {\bfi{Clebsch--Gordan coefficients}.}
831: In an important special case one can relate the branching rules
832: to the so-called Clebsch--Gordan coefficients, which are widely used
833: in physics.
834: 
835: Let us explain 
836: the Clebsch--Gordan coefficients for $su(2)$. Given two
837: representation $D_l$ and $D_k$ (see Section \ref{s.su2rep}), 
838: we can form the tensor 
839: product $D_k\otimes D_l$. An $su(2)$-element $x$ acts on $v\otimes w$ 
840: by mapping 
841: $v\otimes w$ to $x(v)\otimes w + v\otimes x(w)$, where we 
842: write $x(v)$ for the action of $x$ on $v$. In general, the
843: representation $D_k\otimes D_l$ is not irreducible, and  we have
844: \[
845: D_k\otimes D_l
846: = D_{k+l}\oplus D_{k+l-2}\oplus \ldots \oplus D_{|k-l|}\,.
847: \]
848: The precise decomposition of a vector $v\otimes w$, where $v$ and $w$
849: are eigenvectors of $\sigma_3$, into the
850: irreducible components is given by the Clebsch--Gordan
851: coefficients. For the vector in the $D_k$ representation with
852: $\sigma_3$ eigenvalue $m$ and norm one we write $|k,m\rangle$. If the 
853: representation $D_K$ is inside the tensor product of $D_{k_1}$ and
854: $D_{k_2}$, we can decompose any vector $|K, M\rangle$ as a sum of
855: vectors of the form $|k_1,m_1\rangle\otimes |k_2,m_2\rangle$ and the
856: Clebsch--Gordan coefficients are then the coefficients in the
857: decomposition
858: \[
859: |K,M\rangle = \sum_{k_1,k_2m_1,m_2} C_{k_1k_2m_1m_2}^{KM}|k_1,m_1
860: \rangle \otimes |k_2,m_2\rangle \,.
861: \]
862: More generally, the Clebsch--Gordan decompositions say how the tensor
863: product of two irreducible representations $V_1$ and $V_2$ of a
864: compact Lie group (or its Lie algebra) decompose into irreducible
865: representations $W_i$ as $V_1\otimes V_2=\oplus_j W_j$. The
866: Clebsch--Gordan coefficients are the numerical coefficients in the
867: projection from one of the summands $W_j$ to $V_1\otimes V_2$.
868: 
869: 
870: Now suppose that $\Lz_0=\Lz'\oplus \Lz'$ decomposes into two copies 
871: of the same Lie algebra $\Lz'$. An important choice of $\Lz_1$ is
872: the diagonal Lie subalgebra given by elements of the form $(a,a)$ with
873: $a\in  \Lz'$. Then as a Lie algebra $\Lz_1\cong \Lz'$. 
874: The irreducible representations of $\Lz_0$ are given by
875: tensor products of representations of $\Lz_1$. Therefore, decomposing
876: the irreducible representations of $\Lz_0$ with respect to
877: $\tilde\Lz_1$ amounts to giving the Clebsch--Gordan
878: decompositions.
879: 
880: Much more could be said on the topic of symmetries and broken 
881: symmetries in physics. A nice overview is given in a paper by
882: \sca{Bijker} \cite{Bij}. It shows how the symmetry concept organizes
883: not only the world of atoms and molecules that we considered here,
884: but also that of elementary particles. The isospin symmetry between 
885: protons and neutrons has a symmetry group $SU(2)$, which extends to the
886: flavor symmetry group $SU(3)$ for the three light quarks. 
887: 
888: Quantum field theory, culminating in the standard model,
889: is also based on symmetries, namely the space-time symmetries of 
890: the Poincar\'e group, and a gauge group $U(2)\otimes SU(3)$ which
891: combines the broken symmetry group $U(2)=U(1)\otimes SU(2)$ of the weak 
892: interaction (of which only a diagonal subgroup $U(1)$ encoding
893: the electromagnetic charge is unbroken) with the unbroken color 
894: symmetry group $U(3)$ of the strong interaction. 
895: 
896: While these topics lie far beyond the scope of this book,  
897: the interested reader will take the next step and consult deeper 
898: work of others who studies this in depth. Our journey is finished.
899: