cond-mat0605199/so.tex
1: \documentclass[prb,preprint,groupedaddress,showpacs]{revtex4}
2: %\documentclass[prb,twocolumn,groupedaddress,showpacs]{revtex4}
3: \usepackage{graphicx} 
4: \begin{document} 
5: \bibliographystyle{prbrev}
6:  
7: \title{Spin-orbit effects in GaAs quantum wells
8: Spin-orbit effects in GaAs quantum wells:  Interplay
9: between Rashba, Dresselhaus, and Zeeman interactions
10: }
11: 
12: 
13: \author{Enrico Lipparini} 
14: \altaffiliation{Permanent address:
15: Dipartimento di Fisica, Universit\`a di Trento, and INFN,
16: 38050 Povo, Trento, Italy}
17: \affiliation{Departament ECM, Facultat de F\'{\i}sica, 
18: and IN$^2$UB,
19: Universitat de Barcelona. Diagonal 647,
20: 08028 Barcelona, Spain} 
21: 
22: \author{Manuel Barranco} 
23: \affiliation{Departament ECM, Facultat de F\'{\i}sica, 
24: and IN$^2$UB,
25: Universitat de Barcelona. Diagonal 647,
26: 08028 Barcelona, Spain}
27: 
28: \author{Francesc Malet} 
29: \affiliation{Departament ECM, Facultat de F\'{\i}sica, 
30: and IN$^2$UB,
31: Universitat de Barcelona. 
32: Diagonal 647, 08028 Barcelona, Spain}
33: 
34: \author{Mart\'{\i} Pi} 
35: \affiliation{Departament ECM, Facultat de F\'{\i}sica, 
36: and IN$^2$UB,
37: Universitat de Barcelona. Diagonal 647,
38: 08028 Barcelona, Spain}
39: 
40: \author{Lloren\c{c} Serra}
41: \affiliation{Departament de F\'{\i}sica, Universitat de les Illes Balears,
42: and Institut Mediterrani d'Estudis Avan\c{c}ats IMEDEA (CSIC-UIB),
43:  E-07122 Palma de Mallorca, Spain}
44: 
45: 
46: \date{\today}
47: 
48: \begin{abstract} 
49: The interplay between Rashba, Dresselhaus and Zeeman interactions 
50: in a quantum well submitted to an external magnetic field is
51: studied by means of an accurate analytical solution of
52: the Hamiltonian, including electron-electron interactions in a sum rule
53: approach. This solution allows to discuss
54: the influence of the spin-orbit coupling on some relevant
55: quantities that have been measured in inelastic light scattering
56: and electron-spin resonance experiments on  quantum wells.
57: In particular, we have evaluated the spin-orbit contribution to the spin splitting
58: of the Landau levels and to the splitting of charge- and spin-density
59: excitations. We also discuss how the spin-orbit effects change
60: if the applied magnetic field
61: is tilted with respect to the direction perpendicular
62: to the quantum well.
63: \end{abstract}
64: 
65: 
66: \pacs{73.21.Fg, 73.22.Dj, 73.22.Lp}
67: 
68: \maketitle
69: 
70: \section{Introduction}
71: 
72: The study of spin-orbit (SO) effects in semiconductor nanostructures has 
73: been the object of many experimental and 
74: theoretical investigations in the last few years, see e.g. 
75: Refs. \onlinecite{Can99,Ric99,And99,Vos00,Mal00,Rac97, 
76: Vos01,Fol01,Hal01,Ale01,Val02,Val202,Val302,Sch03,Kon05,Cal05} and Refs. 
77: therein. In spite of this,
78:  the extraction from measurements of the effective spin-orbit
79: coupling constant of both Dresselhaus\cite{Dre55} and 
80: Bychkov-Rashba\cite{Ras84,Pik95} SO interactions
81: is not a simple matter, since the SO corrections to the electron
82: energy spectrum in a magnetic field ($B$) are vanishingly small because
83: they correspond to second order effects in perturbation theory. 
84: Thus, few physical observables are sensitive enough to the
85: SO interactions and allow for
86: a quantitative estimate of their coupling constants.
87: One such observable is
88: the splitting of the cyclotron resonance (CR), which has been determined
89: in transmission experiments with far-infrared radiation,\cite{Man01}
90: and is due to the coupling between charge-density
91: and spin-density excitations.\cite{Ton04}
92: A less clear example is the  change in the Larmor frequency 
93: -spin splitting.\cite{Mal06} The spin splitting  has been observed
94: in electron-spin resonance\cite{Ste82,Dob88} and in inelastic
95: light scattering experiments.\cite{Dav97,Kan00}
96: 
97: In this work we extend our previous results\cite{Ton04,Mal06}
98: by obtaining an approximate, yet very accurate, analytical solution
99: of the quantum well SO Hamiltonian that contains both  Dresselhaus and
100: Bychkov-Rashba interactions. In the limit of 
101: high magnetic field, this solution
102: coincides with the results of second order perturbation theory,
103:  and allows to study the SO corrections to 
104: the Landau levels in a simple way, and to study the transitions
105: induced by an external electromagnetic
106: field acting upon the system.
107: 
108: This work is organized as follows. In Sec. II we present the general formalism
109: for the single-particle (sp) Hamiltonian.
110: These results are used in Sec. III to study the transitions 
111: caused by an external electromagnetic field.
112: The role of the electron-electron (e-e) interaction is discussed In Sec.
113: IV
114: within a sum rule approach. In Sec. V we discuss
115: the splitting of the Landau levels  and the appearance of charge- and
116: spin-density modes making, whenever possible, qualitative comparisons
117: with the experimental results.\cite{Man01,Dob88,Eri99,Sal01,Sih04}
118: A brief summary is presented in Sec. VI, and the generalization of
119: some of the expressions derived in Sec. II to the case of tilted
120: magnetic fields is presented in the Appendix.
121: 
122: \section{Single-particle states}
123: 
124: In the effective mass, dielectric constant approximation, the quantum well 
125: Hamiltonian $H$ can be written as $H=H_0
126: + {e^2\over\epsilon}\sum_{i<j=1}^N{1\over\vert{\bf r}_i-{\bf r}_j\vert}$,
127: where $H_0$ is the one-body Hamiltonian  
128: consisting of the 
129: kinetic, Zeeman, Rashba and Dresselhaus 
130: terms
131: %
132: \begin{eqnarray} 
133: H_0 \equiv \sum_{j=1}^N [h_0]_j &=& \sum_{j=1}^N\left[\frac{P^+P^- + P^-P^+}{4m}+ 
134: \frac{1}{2}g^* \mu_B B \sigma_z \right. \nonumber\\ 
135: &+&
136: \left. \frac{\lambda_R}{2 i\hbar}(P^+\sigma_- - 
137: P^-\sigma_+)  
138: + \frac{\lambda_D}{2\hbar}(P^+\sigma_+ + P^-\sigma_-)\right]_j
139: \; .
140: \label{eq1}
141: \end{eqnarray} 
142: %
143: $m=m^* m_e$ is 
144: the effective electron mass in units of the bare electron mass $m_e$, 
145: $P^{\pm}=P_x \pm i P_y$, 
146: $\sigma_{\pm}=\sigma_x \pm i\sigma_y$,
147: where the $\sigma$'s are the Pauli matrices, and
148: ${\bf P}=-i\hbar\nabla+\frac{e}{c}{\bf A}$ represents the 
149: canonical momentum in terms of the vector potential ${\bf A}$, which in the
150: following we write in the Landau 
151: gauge, ${\bf A}= B (0,x,0)$, with ${\bf B}= \nabla \times{\bf A} = B \hat{\bf z}$.
152: The second term in Eq. (\ref{eq1}) is the Zeeman energy, where
153: $\mu_B = \hbar e/(2 m_e c)$ is the Bohr magneton, and $g^*$ is the effective 
154: gyromagnetic factor. The third and fourth terms are the usual Rashba and
155: Dresselhaus interactions, respectively. Note that 
156: for bulk GaAs, taken here as an example,
157: $g^*=-0.44$, $m^*=0.067$, and the dielectric constant 
158: is $\epsilon=12.4$.
159: To simplify the expressions, in the following we
160: shall use effective atomic units 
161: $\hbar=e^2/\epsilon=m=1$. 
162: 
163: Introducing the operators
164: %
165: \begin{equation}
166: \label{eq2}
167: a^{\pm}={1\over\sqrt{2\omega_c}}P^{\pm}
168: \end{equation}
169: %
170: with $[a^-,a^+]=1$ and $\omega_c=e B/c$ being the cyclotron frequency, the 
171: sp  Hamiltonian $h_0$ 
172: can be rewritten as
173: %
174: \begin{equation} 
175: h_0/\omega_c = \frac{1}{2}(a^+a^- + a^-a^+) - {1\over2}{\omega_L\over\omega_c}\sigma_z
176: -\frac{1}{2}\, i\tilde{\lambda}_R(a^+\sigma_- -
177: a^-\sigma_+) 
178: + \frac{1}{2}\, \tilde{\lambda}_D(a^+\sigma_+ + a^-\sigma_-) \;\;\;\; ,
179: \label{eq3}
180: \end{equation} 
181: %
182: where $\omega_L=|g^* \mu_B B|$ is the Larmor frequency
183: and $\tilde{\lambda}_{R,D}=\lambda_{R,D}\sqrt{{2\over\omega_c}}$.
184: For the spinor $|\phi \rangle \equiv \left(\begin{array}{c}
185: \stackrel{\textstyle \phi_1}{\textstyle \phi_2} \end{array} \right)$
186: (we shall  use `1' for the top component and `2' for the bottom component of any 
187: spinor), the Schr\"odinger equation $h_0 |\phi\rangle =\varepsilon |\phi \rangle$ 
188: adopts the form
189: %
190: \begin{equation}
191: \left[ \begin{array}{cc}
192: \frac{1}{2}(a^+a^- + a^-a^+) - \omega_L/(2\omega_c)-\varepsilon &
193: i\tilde{\lambda}_R a^- + \tilde{\lambda}_D a^+ \\
194: -i\tilde{\lambda}_R a^+   +\tilde{\lambda}_D a^- & 
195: \frac{1}{2}(a^+a^- + a^-a^+) + \omega_L/(2\omega_c)-\varepsilon
196:  \end{array}\right]\left(\begin{array}{c}\phi_1\\\phi_2\end{array} \right)=0
197: \;\:\;\; .
198: \label{eq5}
199: \end{equation}
200: %
201: We expand $\phi_1$ and $\phi_2$ into oscillator states $\vert n\rangle$ 
202: as $\phi_1=\sum_{n=0}^\infty a_n\vert n\rangle$, 
203: $\phi_2=\sum_{n=0}^\infty b_n\vert n\rangle$ , on which $a^+$ and $a^-$ act in the 
204: usual way, i.e.,
205:  $\frac{1}{2}(a^+a^- + a^-a^+)\vert n\rangle=(n+\frac{1}{2})\vert n\rangle$,
206: $a^+\vert n\rangle=\sqrt{n+1}\vert n+1\rangle$, $a^-\vert n\rangle=\sqrt{n}\vert 
207: n-1\rangle$,  and $a^-\vert 0\rangle=0$. This yields the infinite system of equations
208: %
209: \begin{eqnarray}
210: (n+\alpha-\varepsilon)b_n
211: -i\tilde{\lambda}_R\sqrt{n}a_{n-1}+\tilde{\lambda}_D\sqrt{n+1}a_{n+1}=0
212: \nonumber\\
213: (n+\beta-\varepsilon)a_n
214: +i\tilde{\lambda}_R\sqrt{n+1}b_{n+1}+\tilde{\lambda}_D\sqrt{n}b_{n-1}=0  
215: \label{eq6}
216: \end{eqnarray}  
217: %
218: for $n\ge0$, with $a_{-1}=0$, $b_{-1}=0$, and
219: $\alpha=(1+\omega_L/\omega_c)/2$, $\beta=(1-\omega_L/\omega_c)/2$.
220: 
221: \subsection{Case in which either \mbox{\boldmath $\lambda_R =0$}, or
222: \mbox{\boldmath $\lambda_D =0$}.}
223: 
224: When only the Rashba or Dresselhaus terms are considered,
225: Eqs. (\ref{eq6}) can be exactly solved.\cite{Ras60,Das90,Fal93,Sch03} 
226: For the sake of completeness, we give here the corresponding results.
227: In the $\lambda_D=0$ case, combining Eqs. (\ref{eq6}) one obtains
228: %
229: \begin{eqnarray}
230: \left[(n+\alpha-\varepsilon)
231: (n-1+\beta-\varepsilon)-n\,{\tilde{\lambda}_R}^2\right]
232: \, b_n &=& 0
233: \nonumber\\
234: \left[(n+\alpha-\varepsilon)
235: (n-1+\beta-\varepsilon)-n\,\tilde{\lambda}_R^2
236: \right] a_{n-1} &=& 0 \;\;\;\; ,
237: \label{eq7}
238: \end{eqnarray}
239: %
240: either of which yields the energies
241: %
242: \begin{equation}
243: \label{eq71}   
244: \varepsilon^{\pm}_n=n\pm\sqrt{{1\over4}
245: \left(1+{\omega_L\over\omega_c}\right)^2
246: +{2\over\omega_c}\lambda^2_R \,n} \;\;\;\; .
247: \end{equation} 
248: %
249: One also obtains
250: %
251: \begin{equation}
252: \label{eq8}
253: (n-1+\beta-\varepsilon^{\pm}_n)\,a^{\varepsilon^{\pm}_n}_{n-1}
254: =-i\tilde{\lambda}_R\sqrt{n}\,b^{\varepsilon^{\pm}_n}_{n} \;\;\;\; ,
255: \end{equation} 
256: %
257: which together with the normalization condition 
258: $\vert a^{\varepsilon^{\pm}_n}_{n-1}\vert^2
259: +\vert b^{\varepsilon^{\pm}_n}_{n}\vert^2=1$ 
260: solves exactly the problem
261: [for $n=0$, $a_{-1}=0$, $b_0=1$, and
262: $\varepsilon_0=\frac{1}{2}(1+\omega_L/\omega_c)$].
263: 
264: Eqs. (\ref{eq7}) indicate that in the series expansion of the spinor
265: $|\phi \rangle$, only one $a_i$ and one $b_i$ coefficient appears.
266: Specifically, 
267: %
268: \begin{equation}
269:  |n_d \rangle =
270:  \left(\begin{array}{c}
271: a^{\varepsilon_n^+}_{n-1}\, |n-1 \rangle \\
272: b^{\varepsilon_n^+}_{n} |n\rangle
273:  \end{array} 
274:  \right) \;\; ; \;\;
275:  |n_u\rangle=
276:  \left(\begin{array}{c}
277: a^{\varepsilon_{n+1}^-}_n\, |n \rangle \\
278: b^{\varepsilon_{n+1}^-}_{n+1} |n+1 \rangle
279:  \end{array} 
280:  \right) \;\;\; .
281: \label{eq8c}
282: \end{equation}
283: %
284: In the limit of zero spin-orbit, the spinors
285: $|n_d \rangle$ and $|n_u \rangle$ become
286: $|n\rangle$$\left(\begin{array}{c} \stackrel{\textstyle 0}{1}
287: \end{array} \right)$ and 
288: $|n\rangle$$\left(\begin{array}{c} \stackrel{\textstyle 1}{0}
289: \end{array} \right)$, respectively.
290: The exact expressions for the $a_i$ and $b_i$ coefficients entering
291: Eq. (\ref{eq8c}) are easy to work out. Expressions valid up to
292: $\lambda_{R,D}^2$ order are given in the next subsection.
293: 
294: The $\lambda_R=0$ case can be worked out similarly. One obtains the
295: secular equation
296: %
297: \begin{equation}
298: \label{eq9}
299: (n+\beta-\varepsilon)(n-1+\alpha-\varepsilon)-n\,{\tilde{\lambda}_D}^2=0
300: \end{equation}
301: %
302: which yields
303: %
304: \begin{equation}
305: \label{eq91}
306: \varepsilon^{\pm}_n=n\pm\sqrt{{1\over4}
307: \left(1-{\omega_L\over\omega_c}\right)^2 +{2\over\omega_c}\lambda^2_D
308: \,n} \;\;\;\; .
309: \end{equation}
310: %
311: One also obtains
312: %
313: \begin{equation}
314: \label{eq10}
315: (n-1+\alpha-\varepsilon^{\pm}_n)\,b^{\varepsilon^{\pm}_n}_{n-1}
316: =-\tilde{\lambda}_D\sqrt{n}\;a^{\varepsilon^{\pm}_n}_{n} \;\;\;\; ,
317: \end{equation}
318: %
319: which together with the normalization condition
320: $\vert a^{\varepsilon^{\pm}_n}_{n}\vert^2
321: +\vert b^{\varepsilon^{\pm}_n}_{n-1}\vert^2=1$ solves exactly the
322: problem (in this case, for $n=0$, $b_{-1}=0$ and $a_0=1$).
323: 
324: Again, in the series expansion of the spinor
325: $|\phi \rangle$, only one $a_i$ and one $b_i$ coefficient appears:
326: %
327: \begin{equation}
328:  |n_d \rangle =
329:  \left(\begin{array}{c}
330: a^{\varepsilon_{n+1}^-}_{n+1}\, |n+1 \rangle \\
331: b^{\varepsilon_{n+1}^-}_{n} |n\rangle
332:  \end{array} 
333:  \right) \;\; ; \;\;
334:  |n_u \rangle =
335:  \left(\begin{array}{c}
336: a^{\varepsilon_n^+}_n\, |n \rangle \\
337: b^{\varepsilon_n^+}_{n-1} |n-1 \rangle
338:  \end{array} 
339:  \right) \;\;\; ,
340: \label{eq10c}
341: \end{equation}
342: %
343: and the same comments as before apply.
344: 
345: \subsection{General case when \mbox{\boldmath $\lambda_R \neq0$} and
346: \mbox{\boldmath $\lambda_D \neq 0$}.}
347: 
348: If both terms are simultaneously considered, the SO interaction
349: couples  the states of all Landau levels, and an exact analytical
350: solution to Eqs. (\ref{eq6}) is unknown, and likely does not exist. 
351: We are going to find an approximate solution that
352: in the  $\lambda^2_{R,D}/\omega_c\ll1$ limit coincides
353: with the results of second order perturbation theory,
354: i.e., it is valid up to $\tilde{\lambda}_{R,D}^2$ order, and it is
355: quite accurate as compared with exact results obtained numerically.
356: Combining Eqs. (\ref{eq6}), one can write
357: %
358: \begin{eqnarray}
359: \label{eq12}
360: \left[n+\alpha-\varepsilon-{\tilde{\lambda}_R}^2{n\over 
361: n-1+\beta-\varepsilon}-{\tilde{\lambda}_D}^2{n+1\over 
362: n+1+\beta-\varepsilon}\right]b_{n}=
363: \nonumber\\
364: \nonumber\\
365: -i\tilde{\lambda}_R\tilde{\lambda}_D\left[{\sqrt{n(n-1)}
366: \over n-1+\beta-\varepsilon}\,b_{n-2}-{\sqrt{(n+1)(n+2)}
367: \over n+1+\beta-\varepsilon}\,b_{n+2}
368: \right]  
369: \end{eqnarray}
370: %
371: and
372: %
373: \begin{eqnarray}
374: \label{eq14}
375: \left[n+\beta-\varepsilon-{\tilde{\lambda}_R}^2{n+1\over
376: n+1+\alpha-\varepsilon}-{\tilde{\lambda}_D}^2{n\over
377: n-1+\alpha-\varepsilon}\right]a_{n}=
378: \nonumber\\
379: \nonumber\\
380: -i\tilde{\lambda}_R\tilde{\lambda}_D\left[{\sqrt{n(n-1)}
381: \over n-1+\alpha-\varepsilon}\,a_{n-2}-{\sqrt{(n+1)(n+2)}
382: \over n+1+\alpha-\varepsilon}\,a_{n+2}
383: \right] .
384: \end{eqnarray}
385: %
386: The approximate solution is obtained 
387: by taking $a_{n-2}=a_{n+2}=b_{n-2}=b_{n+2}=0$ in the above equations. 
388: This means that for each level $|n\rangle$, the SO interaction is 
389: allowed to couple it only with the $|n-1\rangle$ and $|n+1\rangle$ levels.
390: This solution, which
391: consists of a $|n_d\rangle$ and a $|n_u\rangle$ spinor,
392: is therefore obtained by solving first
393: the secular, cubic equation
394: %
395: \begin{equation}
396: \label{eq15}
397: (n+\alpha-\varepsilon)(n-1+\beta-\varepsilon)(n+1+\beta-\varepsilon)=
398: {\tilde{\lambda}_R}^2 n(n+1+\beta-\varepsilon)
399: +{\tilde{\lambda}_D}^2(n+1)(n-1+\beta-\varepsilon)\;\;\;\; .
400: \end{equation}  
401: %
402: Together with the equations
403: %
404: \begin{eqnarray}
405: (n-1+\beta-\varepsilon)a_{n-1}=
406: -i\tilde{\lambda}_R\sqrt{n}\,b_{n}
407: \nonumber\\
408: (n+1+\beta-\varepsilon)a_{n+1}=
409: -\tilde{\lambda}_D\sqrt{n+1}\,b_{n}
410: \label{eq16}
411: \end{eqnarray}  
412: %
413: and the normalization condition $\vert a_{n-1}\vert^2+\vert a_{n+1}\vert^2
414: +\vert b_{n}\vert^2=1$, they determine the $|n_d\rangle$ solution.
415: The  solution corresponding to the $|n_u\rangle$ spinor
416: is obtained by solving the secular equation
417: %
418: \begin{equation}
419: \label{eq17}
420: (n+\beta-\varepsilon)(n-1+\alpha-\varepsilon)(n+1+\alpha-\varepsilon)=
421: {\tilde{\lambda}_R}^2 (n+1)(n-1+\alpha-\varepsilon)
422: +{\tilde{\lambda}_D}^2n(n+1+\alpha-\varepsilon)\;\;\;\; .
423: \end{equation}
424: %
425: Together with the equations
426: %
427: \begin{eqnarray}
428: (n-1+\alpha-\varepsilon)b_{n-1}=
429: -\tilde{\lambda}_D\sqrt{n}\,a_{n}
430: \nonumber\\   
431: (n+1+\alpha-\varepsilon)b_{n+1}=
432: i\tilde{\lambda}_R\sqrt{n+1}\,a_{n}
433: \label{eq18}
434: \end{eqnarray}
435: %
436: and $\vert a_{n}\vert^2+\vert b_{n-1}\vert^2
437: +\vert b_{n+1}\vert^2=1$, they determine the $|n_d\rangle$ solution.
438: 
439: Since all the estimates available in the literature (see for example Refs.
440: \onlinecite{Ton04,Kon05,Mal06} and Refs. therein) yield
441: $\lambda_{R,D}^2$ values of the order of 10 $\mu$eV, and
442: $\omega_c$ in GaAs is of the order of the meV even at
443: small $B (\sim 1$ T),  it is worth to  examine the above solutions in the  
444: ${\tilde{\lambda}_{R,D}}^2=2\lambda_{R,D}^2/\omega_c\ll1$ limit, in
445: which the secular equations 
446: %for the eigenvalues $\varepsilon$
447: have solutions easy to interpret.
448: 
449: To order ${\tilde{\lambda}_{R,D}}^2$, the relevant solution to Eq.
450: (\ref{eq15}) containing both  SO terms is
451: %
452: \begin{equation}
453: \varepsilon_n^d=n+\alpha + 2n{\lambda^2_{R}\over \omega_c+\omega_L}
454: - 2(n+1){\lambda^2_{D}\over \omega_c-\omega_L} \;\;\;\; ,
455: \label{eq19}
456: \end{equation}
457: %
458: that corresponds to the spinor  $| n_d \rangle$
459: %
460: \begin{equation}
461: |n_d \rangle =
462:  \left(\begin{array}{c}
463: a^{\varepsilon_n^d}_{n-1}\, |n-1 \rangle + a^{\varepsilon_n^d}_{n+1}\, |n+1 \rangle
464:  \\ 
465: b^{\varepsilon_n^d}_{n} |n\rangle
466:  \end{array} 
467:  \right)
468: \label{eq19b}
469: \end{equation}
470: %
471: with coefficients
472: %
473: \begin{eqnarray}
474: a^{\varepsilon_n^d}_{n-1}&=&
475: i\tilde{\lambda}_R\sqrt{n}{\omega_c\over \omega_c+\omega_L}
476: \nonumber\\
477: a^{\varepsilon_n^d}_{n+1}&=&
478: -\tilde{\lambda}_D\sqrt{n+1}{\omega_c\over \omega_c-\omega_L}
479: \nonumber\\
480: b^{\varepsilon_n^d}_{n}&=&1 - {1\over2}{\tilde{\lambda}_R}^2
481: n\left({\omega_c\over \omega_c+\omega_L}\right)^2
482: -{1\over2}{\tilde{\lambda}_D}^2(n+1)\left({\omega_c\over \omega_c-\omega_L}\right)^2
483: \;\;\; .
484: \label{eq20}
485: \end{eqnarray}
486: %
487: In the following, we will
488: refer to this solution as to the 
489: quasi spin-down (qdown) solution, since
490: in the zero spin-orbit coupling limit 
491: $|n_d \rangle$ becomes
492: $|n\rangle$$\left(\begin{array}{c} \stackrel{\textstyle 0}{1}
493: \end{array} \right)$.
494: %
495: Analogously, Eq. (\ref{eq17}) has the solution
496: %
497: \begin{equation}
498: \varepsilon_n^u=n+\beta - 2(n+1){\lambda^2_{R}\over 
499: \omega_c+\omega_L}
500: + 2 n {\lambda^2_{D}\over \omega_c-\omega_L} ~~,
501: \label{eq21}
502: \end{equation}
503: %
504: that corresponds to the spinor  $|n_u \rangle$
505: %
506: \begin{equation}
507: |n_u \rangle =
508:  \left(\begin{array}{c}
509: a^{\varepsilon_n^u}_n \, |n \rangle  \\ 
510: b^{\varepsilon_n^u}_{n-1}\, |n-1 \rangle + b^{\varepsilon_n^u}_{n+1}\, |n+1 \rangle
511:  \end{array} 
512:  \right)
513: \label{eq21b}
514: \end{equation}
515: %
516: with coefficients 
517: %
518: \begin{eqnarray}
519: b^{\varepsilon_n^u}_{n-1}&=&
520: \tilde{\lambda}_D\sqrt{n}{\omega_c\over \omega_c-\omega_L}
521: \nonumber\\
522: b^{\varepsilon_n^u}_{n+1}&=&
523: i\tilde{\lambda}_R\sqrt{n+1}{\omega_c\over \omega_c+\omega_L}
524: \nonumber\\
525: a^{\varepsilon_n^u}_{n}&=&1 - {1\over2}{\tilde{\lambda}_R}^2
526: (n+1)\left({\omega_c\over \omega_c+\omega_L}\right)^2
527: -{1\over2}{\tilde{\lambda}_D}^2 n \left({\omega_c\over \omega_c-\omega_L}\right)^2
528: \;\;\; .
529: \label{eq22}
530: \end{eqnarray}
531: %
532: In the following, we will
533: refer to this solution as to the 
534: quasi spin-up (qup) solution, since
535: in the zero spin-orbit coupling limit 
536: $|n_u \rangle$ becomes
537: $|n\rangle$$\left(\begin{array}{c} \stackrel{\textstyle 1}{0}
538: \end{array} \right)$. When either $\lambda_R$ or $\lambda_D$ are 
539: zero, Eqs. (\ref{eq19b}) and (\ref{eq21b}) reduce to the exact Eqs.
540: (\ref{eq8c}) and (\ref{eq10c}), respectively, and the corresponding
541: $a_i$ and $b_i$
542: coefficients, valid up to order $\lambda_{R,D}$, can be extracted from
543: Eqs. (\ref{eq20}) and (\ref{eq22}).
544: These Eqs. show that $a_n$ and $b_n$ are of order $O(1)$, whereas
545: $a_{n\pm1}$ and $b_{n\pm1}$ are of order $O(\lambda_{R,D})$, and
546: $a_{n\pm2}$ and $b_{n\pm2}$ are of order $O(\lambda_{R,D}^2)$.
547: This shows that the neglected terms in Eqs. (\ref{eq12}) and
548: (\ref{eq14}) are of order $O(\lambda_{R,D}^4)$.
549: 
550: The sp energies obtained from Eqs. (\ref{eq19}) and
551: (\ref{eq21}), valid in the $\lambda^2_{R,D}/\omega_c\ll1$ limit,
552: are
553: %
554: \begin{eqnarray}
555: E_n^d &=& (n+{1\over2})\omega_c+{\omega_L\over2} + 2n\lambda^2_{R}{\omega_c\over 
556: \omega_c+\omega_L}
557: - 2(n+1)\lambda^2_{D}{\omega_c\over \omega_c-\omega_L} ~~
558: \nonumber\\
559: E_n^u &=& (n+{1\over2})\omega_c-{\omega_L\over2}  - 2(n+1)\lambda^2_{R}{\omega_c\over 
560: \omega_c+\omega_L}
561: + 2 n \lambda^2_{D}{\omega_c\over \omega_c-\omega_L} \;\;\;\ .
562: \label{eq23}
563: \end{eqnarray}
564: %
565: Together with the structure of the associated spinors,
566: Eqs. (\ref{eq19b}) and (\ref{eq21b}), this sp energy spectrum
567: constitutes one of the main results of our work.
568: By suitable differences of these energies, one may obtain the 
569: sp transition energies discussed in the next Sec.
570: 
571: The above sp energies
572: coincide with  the ones that can be derived from second order
573: perturbation theory with the standard expression
574: %
575: \begin{equation}
576: \label{eq24}
577: E_n^{(2)}={1\over4}\sum_{m\ne n}{\vert\langle m\vert
578: -i\tilde{\lambda}_R\,\omega_c(a^+\sigma_- - a^-\sigma_+)
579: + \tilde{\lambda}_D\,\omega_c(a^+\sigma_+ + a^-\sigma_-)\vert
580: n\rangle\vert^2\over E_n^{0}-E_m^{0}} \;\;\;\; ,
581: \end{equation}
582: %
583: where $|n\rangle= |n,\uparrow\rangle$,
584: $|n,\downarrow\rangle$
585: are the spin-up and spin-down eigenstates of the sp Hamiltonian
586: $\frac{1}{2}(a^+a^- + a^-a^+)\omega_c - {1\over2}\omega_L\sigma_z$
587: with eigenvalues
588: $E_n^{0}(\uparrow)=(n+{1\over2})\omega_c-{1\over2}\omega_L$,
589: and  $E_n^{0}(\downarrow)=(n+{1\over2})\omega_c+{1\over2}\omega_L$, respectively.
590: 
591: The approximate solution Eq. (\ref{eq23}) is very accurate
592: in the high $B$ limit (see below). It also carries an interesting
593: information in the opposite limit of vanishing $B$. In this limit
594: ($\omega_L,\omega_c\ll\lambda^2_{R,D})$, 
595: Eqs. (\ref{eq15}) and (\ref{eq17}) yield the solutions  
596: %
597: \begin{eqnarray}
598: \label{eq25}
599: E_n^d&=&\sqrt{2\omega_c[n \lambda^2_{R}+(n+1)\lambda^2_{D}]}
600: \nonumber\\
601: \nonumber\\
602: E_n^u&=&\sqrt{2\omega_c[(n+1)\lambda^2_{R}+n \lambda^2_{D}]}
603: \end{eqnarray}
604: %
605: which show that, at $B\simeq0$, to order $\lambda^2_{R,D}$ 
606: the Landau levels are not 
607: split due to the SO interaction, as one might
608: naively infer from Eqs. (\ref{eq23}). Another merit of the
609: approximate solution is that it
610: displays in a transparent way the interplay
611: between the three spin-dependent interactions, namely Zeeman,
612: Rashba and Dresselhaus. 
613: Such interplay has been also discussed in Ref. \onlinecite{Mal06},
614: in relation with the violation of the Larmor theorem due to the SO
615: couplings,
616: and in Ref. \onlinecite{Val06}, where the Zeeman and SO interplay is discussed
617: using  the unitarily transformed Hamiltonian technique.
618: Note also that in GaAs quantum wells, which are the object of
619: application in this paper, due to the sign of $g^*$,
620: the lowest energy level is the qup one at the energy
621: $E_0^u={1\over2}\omega_c-{1\over2}\omega_L - 
622: 2\lambda^2_{R}\,\omega_c/(\omega_c+\omega_L)$,
623: containing the Rashba contribution alone, whereas the following level
624: is the qdown one at the energy 
625: $E_0^d={1\over2}\omega_c+{1\over2}\omega_L -
626: 2\lambda^2_{D}\,\omega_c/(\omega_c-\omega_L)$,
627: containing the Dresselhaus contribution alone. For all the other levels 
628: both SO terms contribute to the level energies.
629: 
630: We have assessed the quality of the above analytical solutions, Eqs.\ (\ref{eq23}),
631: by comparing them with exact numerical results for some particular cases.
632: Indeed, 
633: the exact solution to Eqs.\ (\ref{eq6}) can be obtained in the truncated space
634: spanned by the lower ${\cal N}$ oscillator levels. Mathematically, Eqs.\ (\ref{eq6})
635: are then cast into a linear eigenvalue problem of the type
636: %
637: \begin{equation}
638: \label{exd}
639: {\bf M} 
640: \left(\begin{array}{c}
641: \stackrel{\textstyle {\bf a}}{\textstyle {\bf b}} \end{array} \right)
642: = \varepsilon
643: \left(\begin{array}{c}
644: \stackrel{\textstyle {\bf a}}{\textstyle {\bf b}} \end{array} \right)
645: \;\;\;\; ,
646: \end{equation}
647: %
648: where ${\bf M}$ is a $2{\cal N}\times2{\cal N}$ matrix, while ${\bf a}$ and
649: ${\bf b}$ are column vectors made with the sets of coefficients 
650: $\{a_n, n=0,\dots, {\cal N}-1\}$ and $\{b_n, n=0,\dots, {\cal N}-1\}$, respectively.
651: We have diagonalized  ${\bf M}$ using a large enough ${\cal N}$
652: to ensure good convergence in the lower eigenvalues. Fig.\ \ref{fig1} displays
653: a comparison of numerical (symbols) and analytical (solid lines)
654: energies as a function of the Rashba SO strength, for a fixed 
655: Dresselhaus strength, both in units of $\omega_c$, namely
656: $y_R =\lambda_{R}^2/\omega_c$ and  $y_D = \lambda_{D}^2/\omega_c= 0.01$.
657: The chosen values for $y_D$ and $y_R$ are within the
658: expected range for a GaAs quantum well. For instance, if
659: $m \lambda^2_{R,D}/\hbar^2 \sim 10 \mu$eV and $B \sim 1$T,
660: $(m \lambda^2_{R,D}/\hbar^2)/(\hbar \omega_c)\sim 10^{-2}$.
661: There is an excellent agreement between 
662: analytical and numerical results, differences starting to be visible
663: only for strong Rashba intensities
664: and high Landau bands. Actually, 
665: in Fig.\ \ref{fig1} the largest value of the adimensional ratio
666: between Rashba SO and cyclotron energy  
667: %$y_R =\lambda_{R}^2/\omega_c$ in a.u., is 0.05,
668: $y_R =\lambda_{R}^2/\omega_c$ is 0.05, 
669: small enough to validate the analytical expression. Notice, 
670: however, that for larger $y_R$ values -not shown in the figure- i.e., 
671: for small enough $B$, Eqs.\ (\ref{eq23}) no longer reproduces the
672: numerical results. For GaAs this happens
673: for magnetic fields below 0.1 T.
674: Similarly, Fig. \ref{fig2}  displays
675: a comparison of numerical (symbols) and analytical (solid lines)
676: energies as a function of the Dresselhaus SO strength $y_D$, for a fixed
677: Rashba strength $y_R = \lambda_{D}^2/\omega_c= 0.01$.
678: For every Landau level, both figures show a crossing between
679: the $|n_u\rangle$ state, which is at lower energy for 
680: $y_{R,D} \ll 0.01$ because $g^* <0$,  and the $|n_d\rangle$ state,
681: that eventually lies lower in energy. This crossing is due to the
682: interplay between both SO terms.
683: 
684: 
685: \section{Single-particle level transitions induced by applied 
686: electromagnetic fields.}
687: 
688: We can use the preceding results to study the sp transitions induced
689: in the system by the interaction with a
690: left-circular polarized electromagnetic wave 
691: propagating along the $z$-direction, i.e., perpendicular to
692: the plane of motion of the electrons,
693: whose vector potential is ${\bf A}(t)=2A(\cos\theta \hat i+ \sin\theta
694: \hat j)$, with $\theta=\omega t - q z$. The sp
695: interaction Hamiltonian  ${\bf J}\cdot{\bf A}/c +
696: g^* \mu_B\, {\bf s} \cdot (\nabla \times {\bf A})$, where
697: ${\bf J}=e\,{\bf v}/\sqrt{\epsilon}$, reads
698: %
699: \begin{equation}
700: \label{eq26}
701: h_{int}={e\over c\sqrt{\epsilon}}A\left(v_- e^{i\theta}+v_+ e^{-i\theta}\right)
702: +{1\over2}g^*\mu_B q A\left(\sigma_- e^{i\theta}+\sigma_+ e^{-i\theta}\right)
703: \;\;\;\; , 
704: \end{equation}
705: %
706: where the velocity operator $v_\pm$ is
707: defined as $v_\pm \equiv -i[x\pm i y, H] =
708: P^{\pm}\pm i\lambda_R \sigma_\pm + \lambda_D\sigma_\mp$.
709: 
710: The Hamiltonian $h_{int}$ can be rewritten as
711: %
712: \begin{equation}
713: \label{eq271}
714: h_{int}={e\over c\sqrt{\epsilon}}A\sqrt{2\omega_c}\left(\alpha^-
715: e^{i\theta}+\alpha^+ e^{-i\theta}\right)
716: +{1\over2}g^*\mu_B q A\left(\sigma_- e^{i\theta}+\sigma_+
717: e^{-i\theta}\right) \;\;\;\; ,
718: \end{equation}
719: %
720: where the operators $\alpha^+$ and $\alpha^-$ acting on the spinor
721: $|\phi \rangle$ are
722: %
723: \begin{equation}
724: \alpha^+=\left[ \begin{array}{cc}
725: a^+ & i\tilde{\lambda}_R\\\tilde{\lambda}_D & a^+\end{array} \right]~~,
726: ~~\alpha^-=\left[ \begin{array}{cc}
727: a^- & \tilde{\lambda}_D\\-i\tilde{\lambda}_R & a^-\end{array} \right]~~.
728: \label{eq28}
729: \end{equation}
730: %
731: In the dipole approximation ($q \approx 0$),
732: the charge-density excitation operator is  $v_\pm$. We note
733: that, even in the presence of e-e interactions,
734: this operator satisfies the f-sum rule:
735: %
736: \begin{equation}
737: \label{eq29}
738: \sum_n\langle0\vert x\mp i y\vert n \rangle\langle n\vert -iv_\pm\vert0\rangle=
739: \sum_n\omega_{n0}\vert\langle n\vert x\pm i y\vert0\rangle\vert^2=
740: {1\over2}\langle0\vert[x\mp i y,[H,x\pm i y]]\vert0\rangle=2 N~~,
741: \end{equation}
742: %
743: where  $N$ is the electron number and $\omega_{n0}$ are the excitation
744: energies. 
745: 
746: We consider next several useful examples of sp matrix elements
747: involving the operators $\alpha^+$, which is proportional to
748: $v_+$, and $\sigma_-$, and the qup and qdown sp states of Eqs.
749: (\ref{eq19b}) and (\ref{eq21b}).
750: For the operator $\alpha^+$, we can write in general
751: %
752: \begin{equation}
753: \langle \psi |\alpha^+| \phi \rangle  =
754: \psi^*_1a^+\phi_1 +i\tilde{\lambda}_R\psi^*_1\phi_2 
755: +\tilde{\lambda}_D\psi^*_2\phi_1 +\psi^*_2a^+\phi_2
756: \;\;\;\; ,
757: \label{eq30}
758: \end{equation}
759: %
760: and have to distinguish between
761: qup-qup,  qdown-qdown, qup-qdown, and qdown-qup transitions.
762: The qup-qup
763: and qdown-qdown transitions represent
764: the usual CR, and the
765: qup-qdown and qdown-qup are related to spin-flip transitions.
766: 
767: Let us start with the qup-qup and qdown-qdown  transitions.
768: To the order $\lambda^2_{R,D}$ they
769: are dominated by the  transition $n\rightarrow n+1$ at the energies
770: $E_{n+1}^d-E_{n}^d$
771: and 
772: $E_{n+1}^u-E_{n}^u$
773: with  matrix elements 
774: $|\langle (n+1)_u\vert \alpha^+\vert n_u\rangle| =
775: |\langle (n+1)_d\vert \alpha^+\vert n_d\rangle| =\sqrt{n+1}$.
776: The energy splitting of the cyclotron resonance
777: is 
778: %
779: \begin{equation}
780: \label{eq33}
781: \Delta E_{CR} = \left\vert4\lambda^2_{R}{\omega_c\over
782: \omega_c+\omega_L}
783: -4\lambda^2_{D}{\omega_c\over \omega_c-\omega_L}\right\vert ~~.
784: \end{equation}
785: %
786: The $\alpha^+$ excitation operator also induces
787: a qup-qdown transition with  energy
788: $E_{n}^d-E_{n}^u$
789: and matrix element 
790: $|\langle n_{d}\vert \alpha^+\vert n_{u}\rangle|
791: =\tilde{\lambda}_D\,\omega_L/(\omega_c-\omega_L)$.
792: This is a spin-flip transition.
793: In particular, when $n=0$ it is related to the Larmor
794: resonance at the energy\cite{Mal06}
795: %
796: \begin{equation}
797: \label{eq333}
798: \Delta E_{L} = \omega_L + 2\left(\lambda^2_{R}{\omega_c\over
799: \omega_c+\omega_L}
800: -\lambda^2_{D}{\omega_c\over \omega_c-\omega_L}\right) ~~.
801: \end{equation}
802: %
803: Note that the transition matrix 
804: element 
805: is linear in $\tilde{\lambda}_D$, and
806: that in the presence of the Rashba interaction alone,
807: $\alpha^+$ causes no spin-flip transition.
808: 
809: For the operator $\sigma_-$ one gets 
810: $\langle \psi |\sigma_- |\phi \rangle = 2\psi^*_2\phi_1$.
811: The dominant transition is the spin-flip excitation at energy
812: $E_{n}^d-E_{n}^u$ with
813: matrix element $|\langle n_{d}\vert \sigma_-\vert n_{u}\rangle|=2$.
814: The qup-qup and
815: qdown-qdown cyclotron resonances at energies
816: $E_{n+1}^d-E_{n}^d$ and $E_{n+1}^u-E_{n}^u$
817: are also excited with 
818: strengths  $|\langle (n+1)_u\vert \sigma_-\vert n_u\rangle|=
819: |\langle (n+1)_d\vert \sigma_-\vert n_d\rangle|=
820: 2\tilde{\lambda}_D\sqrt{n+1}\,\omega_c/(\omega_c-\omega_L)$.
821: 
822: Other excitations that deserve some  attention are those
823: induced by the operators
824: $\alpha^+\sigma_\pm$ and  $\alpha^+\sigma_z$. 
825: They are detected in inelastic
826: light scattering experiments as spin dipole resonances.\cite{Eri99}
827: The operator $\alpha^+\sigma_z$ excites the same cyclotron
828: states as $\alpha^+$,  at the energies
829: $E_{n+1}^d-E_{n}^d$ and $E_{n+1}^u-E_{n}^u$, and  with
830: the same transition matrix element $\sqrt{n+1}$. 
831: In contrast, the operator
832: $\alpha^+\sigma_+$ mainly induces the transition from  qdown
833: to qup states at the energy  $E_{n+1}^u-E_{n}^d$,
834: whereas the operator
835: $\alpha^+\sigma_-$ induces the transition from  qup
836: to qdown states at the energy $E_{n+1}^d-E_{n}^u$.
837: The transition matrix elements  are given by
838: $|\langle (n+1)_u\vert \alpha^+\sigma_+\vert n_d\rangle| =
839: |\langle (n+1)_d\vert \alpha^+\sigma_-\vert n_u\rangle|
840: =2\sqrt{n+1}$.
841: We  thus see that the dipole transitions between Landau levels
842: $|n\rangle$ and $|n+1\rangle$ at `unperturbed' energies
843: $E_{n+1}-E_{n}$ are split  by the SO interaction,
844: an effect that under some circumstances may be observed,
845: as will be discussed in Sec. V.
846: 
847: \section{Electron-electron interaction and sum rules}
848: 
849: In this section our aim is to discuss the role played by 
850: the e-e interaction in  the physical
851: processes in which SO effects can be  important
852: and, as a consequence, have a chance 
853: to be  experimentally detected. 
854: Since we have obtained a   spinor
855: basis that includes the SO effects (SO basis), one might
856: use it to diagonalize the 
857: Coulomb interaction. This has been done, for
858: example, in Ref. \onlinecite{Cal05}, 
859: where the spinor basis
860: Eq. (\ref{eq8c})  has been used to
861: study the influence of the Rashba interaction 
862: on the incompressible Laughlin state. 
863: One could also
864: use the SO basis Eqs. (\ref{eq19b}) and (\ref{eq21b})
865: to solve the random-phase-approximation
866: equations,\cite{Kal84} or 
867: to study SO effects on the collective states
868: of the quantum well  
869: in the adiabatic time-dependent
870: local-spin-current density approximation.\cite{Mal06,Ser99}
871: We have chosen a different way to incorporate interaction
872: effects that, while being more approximate,
873: it is accurate enough and allows one to obtain simple
874: analytical expressions for the quantities of interest here.
875: It is  the sum rule approach, which is well suited
876: to address the interplay between SO coupling and the
877: e-e interaction in some relevant excitation processes. 
878:  
879: Let us firstly recall
880: that in the absence of the SO coupling,
881: two important theorems hold for the  quantum well Hamiltonian $H$, 
882: in which the e-e interaction is included.
883: They are the Kohn theorem
884: %
885: \begin{equation}
886: \label{eq38}
887: [H,\sum_j P^+_j]=\omega_c\sum_j P^+_j ~~,
888: \end{equation}
889: %
890: which tells  us that, in photoabsorption
891: experiments on quantum wells,
892: a narrow absorption peak must appear at the cyclotron
893: frequency $\omega=\omega_c$ excited by the cyclotron operator
894: $\sum_j P^+_j$, and the Larmor theorem
895: %
896: \begin{equation}
897: \label{eq39}
898: [H,S_-]=\omega_L S_-~~,
899: \end{equation}
900: %
901: which states that in inelastic light scattering experiments at small
902: transferred momentum, or in electron-spin resonance experiments,
903: a narrow collective state 
904: must be excited by the Larmor operator $S_-=\sum_j \sigma^j_-$
905: at the Larmor frequency.
906: These two modes are not  influenced by the e-e interaction.
907: Things radically change if we include in $H$ the  SO interaction. We then obtain
908: %
909: \begin{equation}
910: \label{eq40}
911: [H,\sum_j P^+_j]=\omega_c\sum_j\left( P^+ + i\lambda_R\sigma_+  
912: +\lambda_D\sigma_-\right)_j
913: \end{equation}
914: %
915: and
916: %
917: \begin{equation}
918: \label{eq41}
919: [H,S_-]=\omega_L S_- +\sum_j\left(2i\lambda_R P^-\sigma_z  
920: +2\lambda_D P^+\sigma_z\right)_j \;\;\;\; ,
921: \end{equation}
922: %
923: which show that the SO interaction couples the cyclotron (dipole and
924: spin
925: dipole) and Larmor modes. We have studied in Sec III the effects of the
926: SO coupling on these excitations in the absence of the Coulomb interaction. 
927: Now, we want to determine whether
928: the presence of both the SO and Coulomb interactions 
929: has an effect on the Larmor and cyclotron frequencies or, on the contrary, the
930: results of the previous Sec. still hold. 
931: With this goal in mind, we introduce the
932: following mixed sum rules\cite{Physrep,Lip04}
933: %
934: \begin{eqnarray}
935: m_k^\pm&=&{1\over2} 
936: \sum_n\omega^k_{n0}\left(\langle0\vert F\vert \phi_n\rangle\langle
937: \phi_n\vert G^{\dagger}\vert 0\rangle
938: \pm \langle0\vert G^{\dagger}\vert \phi_n\rangle\langle \phi_n\vert
939: F\vert 0\rangle\right) \nonumber\\
940: &=&{1\over2}\left(\langle0\vert F(H-E_0)^k G^{\dagger}\vert
941: 0\rangle\pm\langle0 \vert G^{\dagger}(H-E_0)^k F\vert 0\rangle\right)
942: \; ,
943: \label{eq42}
944: \end{eqnarray}
945: %
946: where $|0\rangle$ and $|\phi_n\rangle$ are the exact  gs and 
947: excited states of the full Hamiltonian $H$ (including e-e
948: interactions), and $\omega_{n0}=E_n-E_0$ are the corresponding
949: excitation energies. For $k=0-3$ we obtain
950: %
951: \begin{eqnarray}
952: m_0^-&=&{1\over2}\langle0\vert [F, G^{\dagger}]\vert 0\rangle
953: \nonumber\\
954: m_1^+&=&{1\over2}\langle0\vert [F,[H, G^{\dagger}]]\vert 0\rangle
955: \nonumber\\
956: m_2^-&=&{1\over2}\langle0\vert [[F,H],[H, G^{\dagger}]]\vert 0\rangle
957: \nonumber\\
958: m_3^+&=&{1\over2}\langle0\vert [[F,H],[H,[H, G^{\dagger}]]]\vert
959: 0\rangle 
960: \;\;\;\;\;  .
961: \label{eq43}
962: \end{eqnarray} 
963: %
964: Clearly, the
965: more sum rules are known, the better knowledge of the Hamiltonian spectrum.
966: With the four sum rules of Eq. (\ref{eq43})
967: we can obtain information only on two excited states -see below.
968: Consequently,
969: we will limit the analysis to the cases in which either the Rashba or
970: Dresselhaus SO terms are present because, as one can  see from
971: Eq. (\ref{eq40}) and Eq. (\ref{eq41}) as well, in this case  only
972: two states would then be coupled by the corresponding  SO
973: interaction.
974: 
975: Let us first consider the case where  $F=G=\sum_i P^-_i$, i.e.,
976: $G^{\dagger}$ is the cyclotron operator. Evaluating
977: the commutators in Eqs. (\ref{eq43}) we have, to order
978: $\lambda_{R,D}^2$,
979: %
980: \begin{eqnarray}
981: m^-_0 &=& 2N\omega_c
982: \nonumber\\
983: m^+_1 &=& 2N\omega^2_c
984: \nonumber\\
985: m^-_2 &=&  2N\omega^3_c \left[1-{2\over\omega_c}(\lambda^2_{R} -
986: \lambda^2_{D})\right]
987: \nonumber\\
988: m^+_3 &=& 2N\omega^4_c \left[1-{4\over\omega_c}(\lambda^2_{R} -
989: \lambda^2_{D})+{2\omega_L\over\omega^2_c}(\lambda^2_{R} +
990: \lambda^2_{D})\right] \;\;\;\; .
991: \label{eq44}
992: \end{eqnarray}
993: %
994: To obtain these Eqs.  we have used 
995: that $\sum_i P^-_i\vert 0\rangle=0$ and  have assumed that the gs
996: of the system is fully polarized, i.e.,
997: $\langle 0\vert \sum_i \sigma^i_z\vert 0\rangle=N$.
998: As such,
999: these expressions are useless unless the left-hand side can be
1000: directly evaluated from the definition Eq. (\ref{eq42}), and
1001: this evaluation
1002: yields a closed expression for the excitation energies and transition
1003: matrix elements. This is the case if we consider   either of
1004: the $\lambda_{R,D}$ terms alone, because only two states
1005: are excited by the cyclotron operator $G^{\dagger}= \sum_i P^+_i$
1006: acting on the gs $|0\rangle$.
1007: Dropping e.g. the $\lambda_R$ term, a straightforward calculation yields
1008: %
1009: \begin{eqnarray}
1010: \pi_1 &=& 2N\omega_c\left[1 -
1011: {2\omega_c\over(\omega_c-\omega_L)^2}\lambda^2_{D}\right] \nonumber\\
1012: \pi_2 &=& 2N {2\omega_c^2\over(\omega_c-\omega_L)^2}\lambda^2_{D}
1013: \nonumber\\
1014: \omega_{10} &=& \omega_c + {2\omega_c\over\omega_c-\omega_L}\lambda^2_{D} 
1015: \nonumber\\
1016: \omega_{20} &=& \omega_L - {2\omega_c\over\omega_c-\omega_L}\lambda^2_{D}
1017: \;\;\;\; ,
1018: \label{eq45}
1019: \end{eqnarray}  
1020: %
1021: where $\pi_1$ and  $\pi_2$ are the transition strengths
1022: to the cyclotron 
1023: $\vert \phi_{n_1}\rangle$ and Larmor $\vert \phi_{n_2}\rangle$ states,
1024: $\pi_1=\vert \langle \phi_{n_1}\vert \sum_i P^+_i\vert 0\rangle\vert^2$
1025: and
1026: $\pi_2=\vert \langle \phi_{n_2}\vert \sum_i P^+_i\vert 0\rangle\vert^2$,
1027: and $\omega_{10}$, $\omega_{20}$ are the respective excitation
1028: energies.
1029: This is in full agreement with the results of  Sec. III, and
1030: shows that the e-e interaction does not
1031: affect the frequency and transition strengths of the cyclotron
1032: and Larmor resonances. 
1033: 
1034: The case $\lambda_{D}=0$ can be worked out similarly, and the same
1035: conclusion may be extracted. We recall and stress again the results
1036: obtained in the previous section, namely that when $\lambda_{D}=0$,
1037: the Larmor state $\vert \phi_{n_2}\rangle$ is not excited
1038: by the cyclotron operator $\sum_i P^+_i$ ( $\pi_2$ turns out to be
1039: zero).
1040: Alternatively, all previous calculations could have been carried out
1041: using for $G^{\dagger}$ the Larmor operator, namely,
1042: $F=G=\sum_i\sigma^i_+$.
1043: Assuming again that $\langle 0\vert \sum_i \sigma^i_z\vert 0\rangle=N$,
1044: we obtain the same results and draw the same conclusions as before. This
1045: is a consequence of the structure of Eqs. (\ref{eq40}) and (\ref{eq41}).
1046: 
1047: Using more sum rules, e.g. $m^-_4$ and $m^+_5$, one may
1048: obtain information on other states that can be excited by the 
1049: cyclotron operator $\sum_i P^+_i$. Their consideration 
1050: shows that the e-e interaction does not
1051: affect, to order $\lambda^2_{R,D}$,  neither the cyclotron nor the
1052: Larmor state, whose frequencies are the same as determined in
1053: Sec. III when both the SO terms are included in $H$.
1054: 
1055: When the gs of the system has
1056: both qup and qdown occupied states,\cite{Ton04}
1057: the spin dipole operator $\sum_i P^+_i\sigma^i_z$
1058: entering Eq. (\ref{eq41}) excites a state at an energy
1059: $\omega_c(1+{\cal K})$  -see below, instead of
1060: $\omega_c$ as it
1061: corresponds to the cyclotron (charge dipole) operator
1062: $\sum_i P^+_i$, and the results in Eq. (\ref{eq45}) 
1063: must be corrected for. This effect is not related to
1064: the SO interaction, and appears even in the absence of it.
1065: The spin dipole operator does not commute with
1066: the e-e interaction as the
1067: cyclotron operator $\sum_i P^+_i$ does, and ${\cal K}$
1068: is precisely the contribution to the spin dipole operator
1069: $m^+_1$ sum rule
1070: arising from the e-e interaction when one takes
1071: $F=G=\sum_i P^-_i\sigma^i_z$:
1072: %
1073: \begin{eqnarray}
1074: m_1^{+} &=&
1075: \sum_n\omega_{n0}\vert\langle \phi_n\vert \sum_i P^+_i\sigma^i_z\vert
1076: 0\rangle\vert^2 \nonumber\\
1077: &=&{1\over2}\langle0\vert [\sum_j P^-_j\sigma^j_z,[H, \sum_i
1078: P^+_i\sigma^i_z]]\vert 0\rangle=N \omega_c^2(1+{\cal K}) \;\;\;\; ,
1079: \label{eq46}
1080: \end{eqnarray}
1081: %
1082: where
1083: %
1084: \begin{equation}
1085: \label{eq47}
1086: {\cal K}={1\over 2N \omega_c^2}\langle0\vert \sum_{i<j}{\bf\nabla}^2_{r_{ij}}
1087: V(r_{ij})(\sigma^i_z-\sigma^j_z)^2\vert0\rangle~~.
1088: \end{equation}
1089: %
1090: ${\cal K}$ can be extracted from  inelastic light 
1091: scattering experiments.\cite{Eri99} It turns out to be zero
1092: for fully polarized ground states, and small and negative
1093: -of the order of 10$^{-2}$- otherwise. Similarly, the
1094: spin flip dipole operators $\sum_i P^+_i\sigma^i_\pm$,
1095: whose excitations can be also measured by inelastic light scattering,
1096: do not commute with the e-e interaction, which give rise to
1097: some energy corrections. It turns out that
1098: these corrections are equal for the three spin
1099: dipole operators $\sum_i P^+_i\sigma^i_{z,\pm}$ 
1100: because the value of ${\cal K}$ is the same for all them.
1101: Hence, the energy splittings
1102: among these excitations are not influenced by the e-e interaction,
1103: depending only on the Zeeman  and SO energies as found
1104: and discussed at the end of  Sec III.
1105: 
1106: Finally, we want to comment on the consequences of the
1107: failure of the Kohn  theorem due to the
1108: SO coupling using the $m_1^{+}$ sum rule for 
1109: $F=\sum_i P^-_i\sigma^i_z$  and $G=\sum_i P^-_i$:
1110: %
1111: \begin{eqnarray}
1112: m_1^{+} &=&
1113: \sum_n\omega_{n0}\langle0\vert \sum_i P^-_i\sigma^i_z\vert n\rangle\langle n\vert 
1114: \sum_j P^+_j\vert 0\rangle
1115: \nonumber\\
1116: &=&{1\over2}\langle0\vert [\sum_i P^-_i\sigma^i_z,[H, \sum_j
1117: P^+_j]]\vert 0\rangle=\omega_c\langle 0\vert \sum_i \sigma^i_z\vert 0\rangle
1118: \;\;\;\; .
1119: \label{eq48}
1120: \end{eqnarray}  
1121: %
1122: This sum rule allows to study the interplay between 
1123: charge and spin modes. If we cast it into a sum
1124: over `spin dipole states' $\vert m\rangle$ and another
1125: over `charge dipole states' $\vert \rho\rangle$, we obtain
1126: %
1127: \begin{eqnarray}
1128: m_1^{+} &=&
1129: \sum_\rho\omega_{\rho0}\langle0\vert \sum_i P^-_i\sigma^i_z\vert \rho\rangle\langle 
1130: \rho\vert\sum_j P^+_j\vert 0\rangle
1131: \nonumber\\
1132: &+&\sum_m\omega_{m0}\langle0\vert \sum_i P^-_i\sigma^i_z
1133: \vert m\rangle\langle m\vert\sum_j P^+_j\vert 0\rangle
1134: \;\;\;\; .
1135: \label{eq49}
1136: \end{eqnarray}
1137: %
1138: If there is no SO coupling, Kohn's theorem holds, implying that
1139: $\langle m\vert\sum_i P^+_i\vert 0\rangle=0$. Thus, 
1140: when the spin  gs $2S_z=\langle 0\vert \sum_i \sigma^i_z\vert 0\rangle$ is
1141: not zero [otherwise, $m^+_1=0$ from Eq. (\ref{eq48})], only the density
1142: modes
1143: would contribute to $m_1^{+}$ through the $\rho$-sum in Eq. (\ref{eq49}).
1144: On the contrary, if the SO coupling is taken into account, Kohn's theorem 
1145: is violated and the spin and charge dipole states are coupled
1146: to order $\lambda_{R,D}^2$, with $\langle m\vert\sum_i P^+_i\vert 0\rangle$
1147: being now different from zero. 
1148: 
1149: To be more quantitative, let us assume that only one charge dipole
1150: state, the cyclotron state $\vert \rho\rangle$ at energy
1151: $E_1=\omega_c+O(\lambda^2_{R,D})$,
1152: contributes to the first sum in Eq. (\ref{eq49}), and only one
1153: spin dipole state $\vert m\rangle$, at energy
1154: $E_2=\omega_c(1+{\cal K})+O(\lambda^2_{R,D})$,
1155: contributes  to the second sum, where we have indicated
1156: by $O(\lambda^2_{R,D})$ the SO correction to the cyclotron and spin
1157: dipole energies. Let us define the mixed  strengths
1158: %
1159: \begin{eqnarray}
1160: \pi_1 &=& \langle0\vert \sum_i P^-_i\sigma^i_z\vert \rho\rangle\langle
1161: \rho\vert\sum_j P^+_j\vert 0\rangle
1162:  \nonumber\\
1163: \pi_2 &=& \langle0\vert \sum_i P^-_i\sigma^i_z\vert m\rangle
1164: \langle m\vert\sum_j P^+_j\vert 0\rangle
1165: \;\;\;\; .
1166: \label{eq491}
1167: \end{eqnarray}
1168: %
1169: Evaluating the sum rules $m_0^-$ and $m_1^+$  for
1170: the operators $G=\sum_i P^-_i$ and $F=\sum_i P^-_i\sigma^i_z$, one
1171: easily obtains
1172: %
1173: \begin{eqnarray}
1174: \pi_1 &=& \omega_c\langle 0\vert \sum_i \sigma^i_z\vert 0\rangle{E_2-\omega_c\over E_2-E_1}=
1175: \omega_c\langle 0\vert \sum_i \sigma^i_z\vert 0
1176: \rangle\left(1-{O(\lambda^2_{R,D})\over|\omega_c\,{\cal K}|}\right)\; ,
1177:  \nonumber\\
1178: \pi_2 &=& \omega_c\langle 0\vert \sum_i \sigma^i_z\vert 0\rangle{\omega_c-E_1\over E_2-E_1}=
1179: \omega_c\langle 0\vert \sum_i \sigma^i_z\vert 0\rangle 
1180: {O(\lambda^2_{R,D})\over|\omega_c\,{\cal K}|}
1181: \;\;\;\; .
1182: \label{eq492}
1183: \end{eqnarray}
1184: %
1185: Eqs. (\ref{eq492}) explicitly show that if $\langle 0\vert \sum_i \sigma^i_z\vert
1186: 0\rangle=0$,
1187: or if the SO coupling is neglected,  the mixed
1188: strength $\pi_2$ is zero, and the spin dipole state cannot be excited
1189: in photoabsorption experiments.  
1190: The strength $\pi_2$ is nonzero only at odd filling factors $\nu$
1191: ($\nu=2\pi\ell^2 n_e$, where $\ell=\sqrt{\hbar c/e B}$
1192: is the magnetic length and $n_e$ is the electron density),
1193: for which $2 S_z/N=1/\nu$. Besides, when the system is fully polarized
1194: at $\nu=1$, the operators $\sum_i P^-_i$ and $\sum_i P^-_i\sigma^i_z$
1195: coincide and excite the same mode, so there is no splitting.
1196: The SO corrections $O(\lambda^2)$ can be calculated
1197: by taking into account the occupation of the ground state, either using
1198: the sum rule approach of this Section, or the method of unitarily
1199: transforming the Hamiltonian,
1200: as described in Refs. \onlinecite{Ton04,Val06}. This calculation yields
1201: the energy splitting of the CR we discuss in the next Section.
1202: 
1203: \section{Comparison with experiments and discussion}
1204: 
1205: An actual confrontation of the theoretical results we
1206: have obtained with the experiments is not an easy task because of
1207: the smallness of the SO effects, and because of the way they are
1208: presented in the available literature, which
1209: makes it extremely difficult to carry out
1210: a quantitative analysis of such a subtle effect. Thus, we
1211: have to satisfy ourselves with a semi-quantitative analysis, or to point 
1212: out that these results are compatible with fairly rough estimated
1213: values of the SO coupling constants. We present now three such examples
1214: and a possible way to increase SO effects so that they could
1215: be easier to determine.
1216: 
1217: Using unpolarized far-infrared radiation, Manger et
1218: al.\cite{Man01} have measured the cyclotron resonance in 
1219: GaAs quantum wells at different electron densities. The main finding
1220: of the experiment is a well resolved splitting of the CR for  $\nu$=3, 5, and
1221: 7, and no significant splitting for $\nu=1$ and for even filling factors.
1222: We have seen that the SO interaction
1223: couples charge-density and spin-density excitations yielding
1224: the SO splitting of the CR given in Eq. (\ref{eq33}).
1225: However, this expression, by itself,  is unable to explain the filling
1226: factor dependence of the observed splitting, for which
1227: one has to bear in mind  that the SO coupling between the
1228: $\sum_i P^-_i$ and $\sum_i P^-_i\sigma^i_z$ operators is strongly
1229: enhanced when the spin gs is not zero,
1230: as explicitly shown in Eq. (\ref{eq492}). 
1231: We have also noted that ${\cal K}$ contributes to the splitting.
1232: Eq. (\ref{eq33}) has to be generalized to include these features.
1233: We obtain
1234: %
1235: \begin{equation}
1236: \label{eq50}
1237: \Delta E_{CR} = \left\vert{2S_z\over
1238: N}\,4\left(\lambda^2_{R}{\omega_c\over \omega_c+\omega_L}
1239: -\lambda^2_{D}{\omega_c\over \omega_c-\omega_L}\right) + 
1240: {\cal K}\omega_c\right\vert \;\;\;\; ,
1241: \end{equation}
1242: %
1243: where the factor $2S_z/N$ takes into account the actual sp contents
1244: of the gs. This equation, together with Eq. (\ref{eq492}),
1245: embodies the theoretical explanation of the experimental 
1246: findings.\cite{Man01} In particular,
1247: it gives an appreciable splitting only for odd filling factors, for
1248: which the  spin ground state $S_z$ is not zero.
1249: The analysis of the experimental splittings using the Eq. (\ref{eq50}) 
1250: yields values for the quantity 
1251: $m\vert\lambda^2_{R}-\lambda^2_{D}\vert/\hbar^2$ of about $30\mu$eV,
1252: in agreement with the ones recently used to reproduce the spin splitting
1253: in quantum dots\cite{Kon05} and wells.\cite{Mal06} This is, in our
1254: opinion, one of the most clear evidences of a crucial SO effect
1255: on a physical observable, because its absence would imply that
1256: the physical effect does not show up.
1257: 
1258: The spin splitting of the first three Landau levels 
1259: of a GaAs quantum  well has been measured 
1260: in a magnetoresistivity experiment by Dobers et al.\cite{Dob88}
1261: We have shown in the previous sections that this splitting
1262: is not influenced by the e-e interaction, and that there is no
1263: spin splitting as $B$ goes to zero 
1264: [Eq. (\ref{eq25})]. Both facts are in agreement
1265: with the analysis of the experimental data, and with
1266: previous theoretical considerations\cite{Mal86} about the $B$-dependence
1267: of the gyromagnetic factor $g^*$, whose determination
1268: was the physical motivation of the magnetoresistivity 
1269: experiment presented in Ref. \onlinecite{Dob88}.
1270: These authors have derived a $B$- and $n$- dependent $g^*$ factor
1271: $g^*(B,n)= g^*_0 -c(n+\frac{1}{2})B$,
1272: where $g^*_0$ and $c$ are fitting constants that depend on the
1273: actual quantum well. The possibility of a SO shift was not considered,
1274: and their chosen law for $g^*$ implies that the spin splitting energy
1275: $\Delta E_n$ does depend on the Landau level index $n$ entering in a
1276: $B^2$ term, as  they have $\Delta E_n=|g^*\mu_B B|$. A $B$ dependence in
1277: $g^*$ is crucial to explain the experimental data, and also to
1278: reproduce them theoretically.\cite{Mal06}
1279: 
1280: For the spin splitting of the Landau levels we obtain 
1281: %
1282: \begin{equation}
1283: \label{eq51}
1284: \Delta E_n = \omega_L+2(2n+1)\left(\lambda^2_{R}{\omega_c\over
1285: \omega_c+\omega_L} 
1286: -\lambda^2_{D}{\omega_c\over \omega_c-\omega_L}\right) 
1287: %\;\;\;\;\; ,
1288: \end{equation}
1289: %
1290: -recall that $\omega_L=\vert g^*\mu_B B\vert$-
1291: i.e., a splitting that increases with $n$ because of the SO coupling.
1292: This SO correction has been
1293: worked out for the $n=0$ level in Ref. \onlinecite{Mal06} using the
1294: equation of motion method. It is known that the experimental
1295: results\cite{Dob88} for $n=1$ and 2 can be reproduced
1296: if $g^*$ depends on $n$ and $B$, as already shown in that
1297: reference. We have verified that the $n$-dependence of $g^*$ cannot be
1298: mimicked by the $n$-dependence introduced by the SO interaction,
1299: Eq. (\ref{eq51}).
1300: Recently, the analysis of $g^*$ has been extended to
1301: a wider magnetic field range using time-resolved Faraday rotation 
1302: spectroscopy.\cite{Sal01,Sih04}
1303: 
1304: As a third example, we address the inelastic light scattering 
1305: excitation of the spin dipole  modes at $\nu=2$
1306: as measured in Ref. \onlinecite{Eri99}.
1307: For this filling factor, in the absence of SO coupling
1308: the spin-density inter-Landau level 
1309: spectrum is expected to be a triplet mode\cite{Kal84,Lip99}
1310: excited by the three operators $\sum_i P^+_i\sigma^i_{z,\pm}$
1311: with energy splittings
1312: given by the Zeeman energy $\omega_L$. In the presence of SO interactions,
1313: we still expect a triplet mode to appear. Indeed, for $\nu=2$ we have $S_z=0$ and
1314: the cyclotron and  spin dipole
1315: modes excited by the operator $\sum_i P^+_i\sigma^i_z$ are decoupled, as
1316: previously discussed. Thus,
1317: for this operator only one  single mode should be detected at an average
1318: energy $\omega=\omega_c(1+{\cal K})$.
1319: The other operators $\sum_i P^+_i\sigma^i_{\pm}$,
1320: yield the two other spin dipole modes at the energies
1321: %
1322: \begin{equation}
1323: \label{eq52}
1324: E^\pm =\omega_c(1+{\cal K}) \pm \omega_L \pm 4\left(\lambda^2_{R}{\omega_c\over
1325: \omega_c+\omega_L}
1326: -\lambda^2_{D}{\omega_c\over \omega_c-\omega_L}\right) \;\;\;\; .
1327: \end{equation}
1328: %
1329: The splitting is thus symmetric and depends on the
1330: SO strengths. In the experiment, triplet excitations were observed
1331: in all measured samples up to electron densities corresponding to
1332: $r_s=3.3$ (we recall that $r_s=1/\sqrt{\pi n_e}$).
1333: $B$ was accordingly changed to keep the filling factor at $\nu=2$.
1334: Only one triplet mode spectrum at
1335: $B=2.2$ T was shown. From this spectrum, we infer that
1336: there is space for a $\sim 5-10 \%$ SO effect on the splitting, 
1337: assuming that at this fairly small magnetic field,
1338: $g^*$ is that of bulk  GaAs, $g^*=-0.44$. We have estimated that
1339: $m\vert\lambda^2_{R}-\lambda^2_{D}\vert/\hbar^2 \simeq10\mu eV$,
1340: in line with the previous findings. 
1341: Systematic measurements, especially at high $B$
1342: where the splitting is larger, are called for to allow for a
1343: more quantitative analysis.
1344: 
1345: Another unequivocal signature of spin-orbit effects in quantum wells 
1346: would be the detection of the Larmor state in photoabsorption 
1347: experiments. The strength of this transition is given by
1348: $2{\lambda^2_D\over\omega_c}({\omega_L\over\omega_c-\omega_L})^2$ 
1349: and only depends on the Dresselhaus SO coupling 
1350: -see the comment immediately after Eq. (\ref{eq33}).
1351: In most experiments, $B$ is perpendicularly applied
1352: to the plane of motion of the electrons, and for GaAs
1353: the strength is so small that it has never been resolved.
1354: 
1355: We finally discuss the effect of tilting the applied magnetic field
1356: using the expressions derived in the Appendix.
1357: Eq. (\ref{eqa3}) can be used to obtain the
1358: splitting of the cyclotron resonance which generalizes
1359: Eq. (\ref{eq33}) for tilted magnetic fields:
1360: \begin{equation}
1361: \Delta E_{CR}=
1362: 4\left[ (C_R~{\cal V} +C_D~{\cal Z})
1363: \frac{1}{1 + |g^*| m^* {\cal S}/2} -
1364: (C_R~{\cal Z}+C_D~{\cal V})
1365: \frac{1}{1 - |g^*| m^* {\cal S}/2}
1366: \right] \;\;\;\; ,
1367: \label{eq56}
1368: \end{equation}
1369: %
1370: where   $C_{R,D}\equiv m\lambda^2_{R,D}/\hbar^2$,
1371: and the tilting angle $\theta$ enters the quantities ${\cal V}$,
1372: ${\cal Z}$, and ${\cal S}$ defined in the Appendix.
1373: Tilting effects might arise because of the
1374: $1 - |g^*| m^* {\cal S}/2$ denominator in the above equation, but
1375: sizeable effects on $\Delta E_{CR}$ should only be expected for
1376: materials such that  $|g^*| m^*/2$ is large. 
1377: This is not the case for GaAs, but it is, e.g., for
1378: InAs and InSb, which have $|g^*| m^*/2=0.169$ and 0.355, respectively.
1379: For the latter case the dependence of $\Delta E_{CR}$ with 
1380: the in-of-well field $B_x$, with a fixed $B_z$, is shown in Fig.\ 3. 
1381: Notice that $\Delta E_{CR}$ 
1382: is sharply increased when $B_x$ exceeds 
1383: a given value (1T for the parameters in Fig.\ 3), which is proving 
1384: the strong enhancement of SO effects introduced by the horizontal component 
1385: of the tilted field configuration. 
1386: Figure 3 also shows
1387: the comparison with the exact diagonalization data (symbols), indicating that 
1388: the analytical formula, Eq.\ (\ref{eq56}), is accurate up to rather large 
1389: tilting angles and for varying relative weights of Rashba and Dresselhaus terms. 
1390: As a matter of fact, this analytical result does not depend on $B_z$
1391: although, for the sake of comparison with the exact diagonalization, we have used 
1392: $B_z=1$ T in Fig.\ 3. The evolution with $B_x$ is not always monotonous, 
1393: especially
1394: for $C_R>C_D$ where we find an initial decrease of
1395: $\Delta E_{CR}$ with increasing $B_x$, vanishing at $B_x\sim 0.8$ T, 
1396: and eventually increasing again.
1397: 
1398: The tilting also affects the spin splitting of the Landau 
1399: levels 
1400: %
1401: \begin{equation}
1402: {\Delta E_n\over\omega_c}=
1403: \frac{|g^*| m^*}{2} {\cal S} +
1404: 2(2 n + 1)\left[ (y_R~{\cal V} +y_D~{\cal Z})
1405: \frac{1}{1 + |g^*| m^* {\cal S}/2} -
1406: (y_R~{\cal Z}+y_D~{\cal V})
1407: \frac{1}{1 - |g^*| m^* {\cal S}/2}
1408: \right] \;\; , 
1409: \label{eq57}
1410: \end{equation}
1411: %
1412: which generalizes Eq. (\ref{eq51})  for $\theta\ne0$. 
1413: As we have commented before, in a recent experiment
1414: where spin precession frequencies in a InGaAs quantum well
1415:  have been measured 
1416: using electrically detected electron-spin resonances,\cite{Sih04}
1417: a strong dependence of the effective gyromagnetic factor $g^{eff}$
1418: on the applied tilted $B$  has been found. In particular, at
1419: $\theta=45^o$  $g^{eff}$ exhibits oscillations with $B$ which indicate
1420: its sensitivity to the Landau level filling, and a coupling
1421: between spin and orbital eigenstates
1422: which is explicitly present
1423: in the spin-orbit term of Eq.(\ref{eq57}). The effective
1424: g-factor
1425: that can be extracted from this equation at $\theta=45^o$, 
1426: by taking the ratio
1427: $2\Delta E_n/(m^* {\cal S}\omega_c)$,
1428: has the structure 
1429: %
1430: \begin{equation}
1431: \vert g^{eff}(B,n)\vert= \vert g^*_0\vert + \left(n+\frac{1}{2}\right)
1432: \left[c_1 B + {c_2\over B}\right] \;\;\;\; ,
1433: \label{eq58}
1434: \end{equation}
1435: %
1436: where the parametrization $g^*= g^*_0 -c_1(n+\frac{1}{2})B$ of
1437: Refs. \onlinecite{Dob88,Sih04} has been introduced in Eq. (\ref{eq57}),
1438: and the $c_2$ term is the SO contribution.
1439: For the smaller $B$ values in the experiment,
1440: and for reasonable values of  $m\lambda^2_{R,D}/\hbar^2$,
1441: of the order of 1-10$\mu$eV, the SO contribution is 
1442: important enough and should not be
1443: neglected; under these circumstances, time-resolved Faraday
1444: rotation spectroscopy could be sensible to Rashba and/or Dresselhaus
1445: spin-orbit effects.
1446: 
1447: \section{Summary}
1448: 
1449: We have discussed the appearance of spin-orbit effects in
1450: magnetoresistivity and inelastic light scattering experiments on
1451: quantum wells. In particular, we have addressed SO effects on
1452: the splitting of the cyclotron resonance, on the sp Landau level
1453: spectrum, and  on spin-density excitations.
1454: Our discussion has been based on the use of an analytical
1455: solution of the quantum well Hamiltonian valid up to second
1456: order in the SO coupling constants. The accuracy of this
1457: solution has been assessed comparing it with exact numerical
1458: diagonalizations.
1459: 
1460: We have carried out semi-quantitative comparisons with available
1461: experimental data, with the twofold aim of extracting the value
1462: of the SO coupling constants and of indicating possible manifestations
1463: of the SO interactions. We have also pointed out that tilting the
1464: -usually- perpendicularly applied magnetic field might enhance
1465: spin-orbit effects, making them easier to detect.
1466: 
1467: \appendix*
1468: \section{}
1469: 
1470: In this Appendix  we generalize some of the expressions
1471: derived in Sec. II to the  case in which $B$ has a
1472: in-of-well component, e.g., $B=(B_x,0,B_z)$.
1473: The Zeeman term then
1474: becomes $\frac{1}{2} g^* \mu_B {\bf B}\cdot$\mbox{\boldmath $\sigma$}=
1475: $-\frac{1}{2} \omega^z_L (\sigma_x \tan\theta + \sigma_z)$,
1476: where we have introduced  the zenithal angle $\theta$,
1477:  $\tan\theta=B_x/B_z$, and the `$z$-Larmor' frequency
1478: $\omega^z_L=|g^* \mu_B B_z|$, with  $\omega^z_L/\omega_c=|g^*| m^*/2$.
1479: The Schr\"odinger Eq. (\ref{eq5}) then becomes
1480: %
1481: \begin{equation}
1482: \left[ \begin{array}{cc}
1483: \frac{1}{2}(a^+a^- + a^-a^+) - \omega^z_L/(2\omega_c)-\varepsilon
1484: \;\;\;
1485: & i\tilde{\lambda}_R a^- + \tilde{\lambda}_D a^+ 
1486: -[\omega^z_L/(2\omega_c)]\tan\theta  \\
1487: -i\tilde{\lambda}_R a^+   +\tilde{\lambda}_D a^- 
1488: -[\omega^z_L/(2\omega_c)]\tan\theta  \;\;\; &
1489: \frac{1}{2}(a^+a^- + a^-a^+) + \omega^z_L/(2\omega_c)-\varepsilon
1490:  \end{array}\right]\left(\begin{array}{c}\phi_1\\\phi_2\end{array} \right)=0
1491: \;\;\;\; .
1492: \label{eqa1}
1493: \end{equation}
1494: %
1495: The calculation proceeds as before, Eq. (\ref{eq6}) becoming
1496: %
1497: \begin{eqnarray}
1498: (n+\alpha-\varepsilon)~b_n -{\alpha-\beta\over2}\tan\theta~a_n
1499: -i\tilde{\lambda}_R\sqrt{n}~a_{n-1}+\tilde{\lambda}_D\sqrt{n+1}~a_{n+1}=0
1500: \nonumber\\
1501: (n+\beta-\varepsilon)~a_n -{\alpha-\beta\over2}\tan\theta~b_n
1502: +i\tilde{\lambda}_R\sqrt{n+1}~b_{n+1}+\tilde{\lambda}_D\sqrt{n}~b_{n-1}=0  
1503: \;\;\;\; ,
1504: \label{eqa2}
1505: \end{eqnarray}  
1506: %
1507: where $\alpha=(1+\omega^z_L/\omega_c)/2$ and $\beta=(1-\omega^z_L/\omega_c)/2$.
1508: 
1509: The sp spectrum Eq. (\ref{eq23}) becomes
1510: %
1511: \begin{eqnarray}
1512: E_n^d=\left(n+{1\over2}\right)\omega_c+{\omega^z_L\over2}~{\cal S}
1513:  + 2n\left[{\cal U}~(\lambda^2_{R}+ \lambda^2_{D})+
1514: (\lambda^2_{R}~{\cal V}+\lambda^2_{D}~{\cal Z}){\omega_c\over
1515: \omega_c+\omega^z_L~{\cal S}}\right]
1516: \nonumber\\
1517: - 2(n+1)\left[{\cal U}~(\lambda^2_{R}+
1518: \lambda^2_{D})+
1519: (\lambda^2_{R}~{\cal Z}+\lambda^2_{D}~{\cal V}){\omega_c\over
1520: \omega_c-\omega^z_L~{\cal S}}\right]
1521: \nonumber\\
1522: E_n^u=\left(n+{1\over2}\right)\omega_c-{\omega^z_L\over2}~{\cal S} + 
1523: 2n\left[
1524: {\cal U}~(\lambda^2_{R}+ \lambda^2_{D}) +
1525: (\lambda^2_{R}~{\cal Z}+\lambda^2_{D}~{\cal V}){\omega_c\over
1526: \omega_c-\omega^z_L~{\cal S}} 
1527: \right]
1528: \nonumber\\
1529: - 2(n+1)\left[ {\cal U}~(\lambda^2_{R}+ \lambda^2_{D}) +
1530: (\lambda^2_{R}~{\cal V}+\lambda^2_{D}~{\cal Z}){\omega_c\over
1531: \omega_c+\omega^z_L~{\cal S}}
1532: \right]  \;\;\;\; ,
1533: \label{eqa3}
1534: \end{eqnarray}
1535: %
1536: where we have defined ${\cal S}=1/\cos\theta$, ${\cal U}=\sin^2\theta/4$,
1537: ${\cal V}=(1+ \cos\theta)^2/4$, and ${\cal Z}=(1- \cos\theta)^2/4$.
1538: When $\theta=0,\,{\cal U}= {\cal Z}=0,\, {\cal V}=1$, and
1539: Eq. (\ref{eqa3}) reduces to Eq. (\ref{eq23}).
1540: 
1541: \section*{ACKNOWLEDGMENTS} This work has been performed 
1542: under grants FIS2005-01414 and FIS2005-02796 from
1543: DGI (Spain) and 2005SGR00343 from Generalitat de Catalunya.
1544: E. L. has been suported by DGU (Spain), grant SAB2004-0091.
1545: 
1546: 
1547:  
1548: \begin{thebibliography}{99}
1549: 
1550: \bibitem{Can99} C-M. Hu, J. Nitta, T. Akazaki, H. Takayanaga, J. Osaka,
1551: P. Pfeffer, and W. Zawadzki, Phys. Rev. B {\bf60}, 7736 (1999).
1552: 
1553: \bibitem{Ric99} D. Richards and B. Jusserand, Phys. Rev. B {\bf59}, R2506 (1999).
1554: 
1555: \bibitem{And99} E. de Andrada e Silva, Phys. Rev. B {\bf60}, 8859 (1999).
1556: 
1557: \bibitem{Vos00} A. Voskoboynikov, S.S. Liu, C.P. Lee, and O. Tretyak, 
1558: J. App. Phys. {\bf87}, 1 (2000).
1559: 
1560: \bibitem{Mal00} A.G. Mal'shukov and K.A. Chao, Phys. Rev. B {\bf61}, R2413 (2000).
1561: 
1562: \bibitem{Rac97} P.N. Racec, T. Stoica, C. Popescu, M. Lepsa, and T.G. van de Roer, 
1563: Phys. Rev. B {\bf56}, 3595 (1997).
1564: 
1565: \bibitem{Vos01} O. Voskoboynikov, C.P. Lee, and O. Tretyak, Phys.\ Rev.\ B {\bf 63},
1566: 165306 (2001).
1567: 
1568: \bibitem{Fol01} J. A. Folk, S.R. Patel, K.M. Birnbaum, C.M. Marcus, C.I. Duru\"oz,
1569:  and J.S. Harris, Jr, Phys.\ Rev.\ Lett.\ {\bf 86}, 2102 (2001).
1570: 
1571: \bibitem{Hal01} B.I. Halperin, A. Stern, Y. Oreg, J.N.H.J. Cremers, J.A.
1572: Folk, and C.M. Marcus, Phys.\ Rev.\ Lett.\ {\bf 86}, 2106 (2001).
1573: 
1574: \bibitem{Ale01} I.L. Aleiner and V. I. Fal'ko, Phys.\ Rev.\ Lett.\ {\bf
1575: 87}, 256801 (2001).
1576: 
1577: \bibitem{Val02} M. Val\'{\i}n-Rodr\'{\i}guez, A. Puente, Ll.\ Serra, and 
1578: E. Lipparini, Phys. Rev. B {\bf 66}, 165302 
1579: (2002).
1580: 
1581: \bibitem{Val202} M. Val\'{\i}n-Rodr\'{\i}guez, A. Puente, and Ll. Serra, 
1582: Phys. Rev. B {\bf 66}, 045317 (2002).
1583: 
1584: \bibitem{Val302} M. Val\'{\i}n-Rodr\'{\i}guez, A. Puente, Ll.\ Serra, and 
1585: E. Lipparini, Phys. Rev. B {\bf 66}, 235322 
1586: (2002).
1587: 
1588: \bibitem{Sch03} J. Schliemann, J.C. Egues, and D. Loss, Phys. Rev. B {\bf 67},
1589: 085302 (2003).
1590: 
1591: \bibitem{Kon05} J. K\"onemann, R.J. Haug, D.K. Maude, V.I. Fal'ko, and 
1592: B.L. Altshuler, Phys. Rev. Lett. {\bf 94}, 
1593: 226404 (2005).
1594: 
1595: \bibitem{Cal05} M. Califano, T. Chakraborty, P. Pietilainen,
1596: Phys. Rev. Lett. {\bf 94}, 246801 (2005).
1597: 
1598: \bibitem{Dre55} G. Dresselhaus, Phys. Rev. {\bf 100}, 580 (1955).
1599: 
1600: \bibitem{Ras84} Yu. A. Bychkov and E.I. Rashba, J. Phys. C {\bf 17}, 6039 (1984).
1601: 
1602: \bibitem{Pik95} F.G. Pikus and G.E. Pikus, Phys. Rev. 
1603: B {\bf 51}, 16928 (1995).
1604: 
1605: \bibitem{Man01} M. Manger, E. Batke, R. Hey, K.J. Friedland, K. K\"ohler,
1606: and P. Ganser,  Phys. Rev. B{ \bf 63}, 121203(R) (2001).
1607: 
1608: \bibitem{Ton04} P. Tonello and E. Lipparini, Phys. Rev. B{ \bf 70}, 081201(R) (2004).
1609: 
1610: \bibitem{Mal06} F. Malet, E. Lipparini, M. Barranco, and M. Pi, Phys. Rev. B
1611: {\bf 73}, 125302 (2006).
1612: 
1613: \bibitem{Ste82}
1614: D. Stein, K. v. Klitzing, and G. Weimann, Phys. Rev. Lett. {\bf 51},
1615: 130 (1982).
1616: 
1617: \bibitem{Dob88} 
1618: M. Dobers, K. v. Klitzing, and G. Weimann, Phys. Rev. B {\bf 38}, 5453 (1988).
1619: 
1620: \bibitem{Dav97} H.D.M. Davies, J.C. Harris, J.F. Ryan, and A.J. Turberfield, 
1621: Phys. Rev. Lett. {\bf 78}, 4095 (1997).
1622: 
1623: \bibitem{Kan00} M. Kang, A. Pinczuk, B.S. Dennis, M.A. Eriksson, 
1624: L.N. Pfeiffer, and K.W. West, Phys. Rev. Lett. {\bf 84}, 546 
1625: (2000).
1626: 
1627: \bibitem{Eri99} M.A. Eriksson, A. Pinczuk, B.S. Dennis, S.H. Simon, 
1628: L.N. Pfeiffer and K.W. West, Phys. Rev. Lett. {\bf 82 }, 2163 (1999).
1629: 
1630: \bibitem{Sal01} G. Salis, D.D. Awschalom, Y. Ohno, and H. Ohno, 
1631: Phys. Rev. B {\bf 64}, 195304 (2001).
1632: 
1633: \bibitem{Sih04} V. Sih, W.H. Lau, R.C. Myers, A.C. Gossard, M.E.
1634: Flatt\'e, and D.D. Awschalom, Phys. Rev. B {\bf 70}, 161313(R) (2004).
1635: 
1636: \bibitem{Ras60} E.I. Rashba, Fiz. Tverd. Tela (Leningrad) {\bf 2}, 1224 (1960)
1637: [Sov. Phys. Solid State {\bf 2}, 1109 (1960)].
1638: 
1639: \bibitem{Das90} B. Das, S. Datta and R. Reifenberger, Phys. Rev. B {\bf 41}, 8278
1640: (1990).
1641: 
1642: \bibitem{Fal93} V.I. Falko, Phys. Rev. B {\bf 46}, R4320 (1992).
1643: 
1644: \bibitem{Val06} M. Val\'{\i}n-Rodr\'{\i}guez and R. G. Nazmitdinov,
1645: cond-mat/0512231.
1646: 
1647: \bibitem{Kal84} C. Kallin and B.I. Halperin, Phys. Rev. B {\bf 30}, 5655 (1984).
1648: 
1649: \bibitem{Ser99} 
1650: Ll.\ Serra, M. Barranco, A. Emperador, M. Pi, E. Lipparini,
1651: Phys.\ Rev.\ B {\bf 59}, 15290 (1999).
1652: 
1653: \bibitem{Physrep} E. Lipparini and S. Stringari, Phys. Rep.
1654: {\bf 175}, 103 (1989).
1655: 
1656: \bibitem{Lip04} E. Lipparini, 
1657: {\it Modern Many Particle Physics-Atomic Gases, Quantum Dots
1658: and Quantum Fluids} (World Scientific, Singapore 2003).
1659: 
1660: \bibitem{Mal86} F. Malcher, G. Lommer, and U. R\"ossler, Superlattices
1661: and Microstructures {\bf 2}, 267 (1986);
1662: G. Lommer, F. Malcher, and U. R\"ossler, {\it ibid.} {\bf 2}, 273
1663: (1986).
1664: 
1665: \bibitem{Lip99} E. Lipparini, M. Barranco, A. Emperador, M. Pi, and
1666: Ll. Serra, Phys. Rev. B {\bf 60}, 8734 (1999).
1667: 
1668: \end{thebibliography}
1669: 
1670: % FIG.1 
1671: \pagebreak
1672: 
1673: \begin{figure}[t] 
1674: \centerline{\includegraphics[width=14cm,clip]{fig1.eps}}
1675: \caption{
1676: Lower energy levels for a GaAs quantum well as a function of the Rashba
1677: intensity $y_R =\lambda_{R}^2/\omega_c$ 
1678:  for a fixed Dresselhaus intensity
1679: $y_D = \lambda_{D}^2/\omega_c=0.01$.
1680:  Solid
1681: lines are the analytical result, Eq.\ (\ref{eq23}), while
1682: symbols correspond to the exact diagonalization, Eq.\ (\ref{exd}).
1683: } 
1684: \label{fig1} 
1685: \end{figure}
1686: 
1687: % FIG.2
1688: %\pagebreak
1689: 
1690: \begin{figure}[t] 
1691: \centerline{\includegraphics[width=14cm,clip]{fig2.eps}}
1692: \caption{
1693: Lower energy levels for a GaAs quantum well as a function of the
1694: Dresselhaus intensity $y_D = \lambda_{D}^2/\omega_c$ 
1695:  for a fixed  Rashba intensity $y_R =\lambda_{R}^2/\omega_c=0.01$.
1696:   Solid
1697: lines are the analytical result, Eq.\ (\ref{eq23}), while
1698: symbols correspond to the exact diagonalization, Eq.\ (\ref{exd}).
1699: } 
1700: \label{fig2}
1701: \end{figure}
1702: 
1703: % FIG.3
1704: %\pagebreak
1705: 
1706: \begin{figure}[t] 
1707: \centerline{\includegraphics[width=14cm,clip]{fig3.eps}}
1708: \caption{
1709: Splitting of the cyclotron resonance for an InSb quantum well ($|g^*|m^*/2=0.355$) 
1710: as a function of the in-of-well field $B_x$ when $B_z=1$ T.
1711: Lines are the result from the analytical formula, Eq.\ (\ref{eq56}), while symbols correspond
1712: to the exact diagonalization of Eq.\ (\ref{eqa1}).
1713: Defining $C_{R,D}=m\lambda_{R,D}^2/\hbar^2$ the shown results are for:
1714: $C_R=30$ $\mu$eV and $C_D=10$ $\mu$eV, solid line and circles; 
1715: $C_R=10$ $\mu$eV and $C_D=10$ $\mu$eV, long-dashed line and triangles;
1716: $C_R=10$ $\mu$eV and $C_D=30$ $\mu$eV, short-dashed line and squares.
1717: } 
1718: \label{fig3}
1719: \end{figure}
1720: 
1721: 
1722: 
1723: \end{document}
1724: