1: \documentclass[a4paper,pra,showpacs,superscriptaddress,twocolumn]{revtex4}
2:
3: % \documentclass[12pt,a4paper,pra,showpacs,superscriptaddress]{revtex4}
4:
5: \usepackage{amsfonts}
6: \usepackage{amssymb}
7: \usepackage{amsmath}
8: \usepackage{calc}
9: \usepackage{graphicx}
10: \usepackage{array}
11: \usepackage{bm}
12: \usepackage{delarray}
13: \usepackage{ifthen}
14:
15: \newcommand{\sqrtfrac}[2]{\ifthenelse{\equal{#1}{1}}{\frac{\,\,1}{\sqrt{#2}}}{\frac{\sqrt{#1}}{\sqrt{#2}}}}
16:
17: \def\advecb{{\bm\Phi}}
18: \def\advec{{\Phi}}
19: \def\hdots{\cdots}
20: \def\t{(t)}
21: \def\eigenv{\varepsilon}
22:
23: % here are some colors ------------------------
24: \def\Blue{\special{color cmyk 1. 1. 0 0}} % PANTONE BLUE-072
25: \def\Green{\special{color cmyk .6 .3 1. 0}} % MY GREEN
26: \def\Red{\special{color cmyk 0 1. 1. 0}} % PANTONE RED
27: \def\Black{\special{color cmyk 0 0 0 1.}} % PANTONE PROCESS-BLACK
28: \def\Blue{\special{color cmyk 1. 1. 0 0}} % PANTONE BLUE-072
29: \def\Brown{\special{color cmyk 0 0.81 1. 0.60}} % PANTONE 1615
30:
31: \def\bs#1{#1}
32: \def\bwsbox#1{}
33: \def\zk#1{#1}
34: \def\zkbox#1{}
35: %% omit the following when finished debugging
36: \def\bws#1{{\Red #1 \Black}} %% for debugging
37: \def\bwsbox#1{\fbox{BWS: #1}} %% for debugging
38: \def\zk#1{\Green#1\Black} %% for debugging
39: \def\zkbox#1{\fbox{ZK: #1}} %% for debugging
40: \def\nv#1{\Red #1 \Black} %% for debugging
41: \def\nvbox#1{\fbox{NV: #1}} %% for debugging
42:
43: \def\bmz{{\bm z}}
44:
45:
46:
47: %%%%%%%%%%%%%%%%%%%%%%%%%% begin document %%%%%%%%%%%%
48:
49: \begin{document}
50: \title{Stimulated Raman Adiabatic Passage (STIRAP) Among
51: Degenerate-Level Manifolds}
52:
53: \author{Z. Kis}
54: \affiliation{H.A.S. Research Institute for Solid State Physics and
55: Optics, H-1525 Budapest, P.O.Box 49, Hungary}
56: \affiliation{Fachbereich Physik der Universit\"at Kaiserslautern,
57: 67653 Kaiserslautern, Germany}
58: \author{A. Karpati}
59: \affiliation{H.A.S. Research Institute for Solid State Physics and
60: Optics, H-1525 Budapest, P.O.Box 49, Hungary}
61: \author{B. W. Shore}
62: \affiliation{618 Escondido Cir., Livermore CA 94550, USA}
63: \affiliation{Fachbereich Physik der Universit\"at Kaiserslautern,
64: 67653 Kaiserslautern, Germany}
65: \author{N. V. Vitanov}
66: \affiliation{Department of Physics, Sofia University, James Boucher 5
67: blvd., 1164 Sofia, Bulgaria}
68: \altaffiliation{Institute of Solid State Physics,
69: Bulgarian Academy of Sciences, Tsarigradsko chauss\'{e}e 72, 1784
70: Sofia, Bulgaria}
71: \affiliation{Fachbereich Physik der Universit\"at Kaiserslautern,
72: 67653 Kaiserslautern, Germany}
73:
74: \begin{abstract}
75: We examine the conditions needed to accomplish stimulated Raman
76: adiabatic passage (STIRAP) when the three levels ($g$, $e$ and $f$)
77: are degenerate, with arbitrary couplings contributing to the
78: pump-pulse interaction ($g$ - $e$) and to the Stokes-pulse
79: interaction ($e$-$f$). We show that in general a {\em sufficient}
80: condition for complete population removal from the $g$ set of
81: degenerate states for arbitrary, pure or mixed, initial state is
82: that the degeneracies should not decrease along the sequence $g$,
83: $e$ and $f$. We show that when this condition holds it is possible
84: to achieve the degenerate counterpart of conventional STIRAP,
85: whereby adiabatic passage produces complete population transfer.
86: Indeed, the system is equivalent to a set of independent three-state
87: systems, in each of which a STIRAP procedure can be implemented. We
88: describe a scheme of unitary transformations that produces this
89: result. We also examine the cases when this degeneracy
90: constraint does not hold, and show what can be accomplished in those
91: cases. For example, for angular momentum states when the degeneracy
92: of the $g$ and $f$ levels is less than that of the $e$ level we show
93: how a special choice for the pulse polarizations and phases can
94: produce complete removal of population from the $g$ set. Our scheme
95: can be a powerful tool for coherent control in degenerate systems,
96: because of its robustness when selective addressing of the states is
97: not required or impossible. We illustrate the analysis with several
98: analytically solvable examples, in which the degeneracies originate
99: from angular momentum orientation, as expressed by magnetic
100: sublevels.
101: \end{abstract}
102:
103: \date{\today }
104:
105: \pacs{32.80.Qk,42.65.Dr,33.80.Be}
106:
107: \maketitle
108: % \tableofcontents
109:
110: %------------------------------------------------------
111: \section{Introduction}
112: %------------------------------------------------------
113:
114: Techniques based on adiabatic passage provide very practical methods
115: for producing nearly complete transfer of population between two
116: quantum states using crafted laser pulses \cite{Vitanov01}. One
117: popular example of such coherent adiabatic excitation, stimulated
118: Raman adiabatic passage (STIRAP) \cite{STIRAP}, provides a simple and
119: robust technique for transferring population between two nondegenerate
120: metastable levels, making use of two pulses, termed the pump pulse
121: (linking the initially populated ground state $\psi_g$ with excited
122: state $\psi_e$) and the Stokes pulse (linking excited state $\psi_e$
123: with final state $\psi_f$ of the three-state chain). When the pulses
124: are properly timed (Stokes preceding but overlapping the pump pulse)
125: and two-photon resonance is maintained, then via adiabatic passage the
126: population is transferred from initial to final state, without
127: appreciable population in the excited state at any time.
128:
129: The operation of STIRAP can be understood by introducing instantaneous
130: eigenstates of the time-varying Hamiltonian, the time-dependent
131: adiabatic states with associated time-dependent eigenvalues (adiabatic
132: energies). One (and only one) of these states, $\Phi_0(t)$, is
133: constructed from only the initial and final state, with no component
134: of the excited state. Because the excited state generates
135: fluorescence via spontaneous emission, such an adiabatic state will
136: exhibit no such signal; it is termed a {\em dark} state. During the
137: STIRAP process the state vector $\Psi(t)$ remains aligned with the
138: adiabatic state $\Phi_0(t)$, while this state, in turn, changes
139: composition from being aligned with $\psi_g$ initially to being
140: aligned with $\psi_f$ after the Stokes-pump pulse sequence.
141:
142: Numerous extensions of the basic three-state STIRAP \cite{STIRAP} have
143: been considered \cite{Vitanov01,AAMOP}, including examples in which
144: there occur magnetic sublevels and associated degeneracy. One
145: possibility is that the atomic energy levels are coupled in such a way
146: that each one is connected to at most two others. Population transfer
147: in such multi-state chains has been studied by several authors
148: \cite{Shore91, Smith92, Pillet93, Weiss94, Shore95, Martin95,
149: Malinovsky97, Vitanov98, Theuer98}. In addition to straightforward
150: population transfer, STIRAP has been applied to the problem of
151: manipulating and creating coherent superpositions of two or more
152: quantum states. Such superpositions are required for many
153: contemporary applications including information processing and
154: communication. The original STIRAP process has, for example, been
155: utilized to create coherent superpositions in three- and four-level
156: systems \cite{Marte91, Lawall94, Weitz94, Goldner94, Unanyan98,
157: Theuer99, Unanyan99} and to prepare $N$-component maximally coherent
158: superposition states \cite{Unanyan01}. There have been proposals to
159: create $N$-component coherent superpositions in such systems, where
160: the final state space is degenerate \cite{Kis01, Kis02}, at least in
161: the rotating wave picture. This idea has been further developed to
162: map wave-packets between vibrational potential surfaces in molecules
163: \cite{Kraal02a, Kraal02b}. Finally, it has been shown for a specific
164: degenerate system, having a single initial-, two degenerate
165: intermediate-, and three degenerate final states coupled in the Raman
166: configuration, that the STIRAP process can be extended to systems with
167: degenerate intermediate and final levels \cite{Kis03}.
168:
169: Yet an open question has remained: what is the most general system of
170: three degenerate levels, linked via Raman process, for which it is
171: possible to transfer all population from the ground-state manifold of
172: degenerate states (the $g$ set) to the final-state manifold (the $f$
173: set) while minimizing population in the excited states (the $e$ set),
174: without first using optical pumping to prepare a single nondegenerate
175: initial state? We here provide the answer to this question.
176:
177:
178: We consider $N_g$ degenerate states of the $g$ set, coupled by means
179: of a pump-pulse to $N_e$ degenerate states of the $e$ set, which in
180: turn are linked by the Stokes pulse to $N_f$ degenerate states of the
181: $f$ set. We will show that such a generalized STIRAP process is
182: almost always possible if the succession of state-degeneracies is
183: nondecreasing, i.e. $ N_g \leq N_e \leq N_f$. When such conditions
184: hold, then for arbitrary couplings among states (e.g. arbitrary
185: elliptical polarization of electric dipole radiation between magnetic
186: sublevels) it is possible to obtain complete adiabatic passage of all
187: population from the states of the $g$ set into some combination of
188: states of the $f$ set.
189:
190: We also examine the possibility of adiabatic passage when this
191: restriction on degeneracies does not hold. We show that in this case
192: in general only part of the population can be transferred to the $f$
193: set. We point out that, in special but important cases, for an
194: appropriate choice of the polarizations and phases of the coupling
195: fields, a complete adiabatic population transfer can be obtained.
196:
197: Another motivation of this paper is the creation of coherent
198: superposition states in a degenerate system. The difficulty in such
199: systems arises from the limited possibility of addressing a single
200: preselected state: addressing of a selected state is usually achieved
201: by exploiting selection rules that the coupling field should satisfy.
202: However, if we have e.g. two Zeeman multiplets a light field with a
203: certain polarization will create several couplings between the
204: magnetic sublevels of the multiplets. Our scheme offers a solution to
205: this problem: we show that despite of the lack of selective addressing
206: of the degenerate states, we have some control over the created
207: coherent superposition state in the $f$ set. As we point out, and
208: illustrate with specific examples, the level of control depends on the
209: system under consideration.
210:
211: Our scheme is based on using a Morris-Shore (MS) transformation of the
212: Stokes couplings or the pump couplings, thereby reducing this
213: particular (generally complicated) linkage to a set of unlinked
214: two-state systems and dark states \cite{Morris83,5ss}. Underlying
215: this technique is the fact that, as Morris and Shore \cite{Morris83}
216: have shown, any system of linkages in which there occur only two
217: detunings (i.e. the system has two sets of degenerate sublevels,
218: termed here $a$ and $b$, forming sets of dimension $N_a$ and $N_b$),
219: can be transformed, via suitable redefinition of basis states, to one
220: involving a set of $N_{<}$ independent two-state systems, where
221: $N_{<}={\rm min}\{N_a, N_b\}$, together with a set of uncoupled states
222: that are unconnected to other states by the given couplings (one-state
223: systems). If such an uncoupled state has no component from the $e$
224: set we term it a {\em dark} state. We here extend that work to
225: produce sets of unlinked three-state systems.
226:
227:
228: The paper is organized as follows: In the next section we present a
229: general model for degenerate, three-level systems and discuss its main
230: properties. In Sec.~\ref{sec:mstrafo} we derive a general condition
231: for complete STIRAP-like population transfer. In
232: Sec.~\ref{sec:stokes-mstrafo} we derive analytic expressions for the
233: dark and bright states for important special choices of degeneracies.
234: Then, in Sec.~\ref{sec:adiab-tevol}, we determine the conditions
235: needed for adiabatic evolution. We demonstrate our method through
236: some specific examples in Sec.~\ref{sec:examples}. Finally, in
237: Sec.~\ref{sec:summary}, we summarize our results.
238:
239: %------------------------------------------------------
240: \section{The Degenerate-Sublevel Model} \label{sec:model}
241: %------------------------------------------------------
242:
243: \subsection{The Hamiltonian}
244:
245: As is customary when dealing with STIRAP or other three-level chains,
246: we introduce an expansion of the state vector $\Psi(t)$ that
247: incorporates explicit phases taken from carrier frequencies of the
248: pump and Stokes pulses, $\omega_p$ and $\omega_S$, respectively. In
249: this rotating-wave picture, and with the customary neglect of
250: counter-rotating terms [i.e. time variations $(\omega_i +
251: \omega_j)t$] the rotating-wave approximation (RWA) Hamiltonian takes
252: the block-matrix form
253: \begin{equation}\label{ham}
254: {\bm H}(t)=\left[\begin{array}{ccc}
255: {\bm 0}&p(t){\bm P}&{\bm 0}\\
256: p(t){\bm P}^\dagger&\hbar{\bm \Delta}&s(t){\bm S}\\
257: {\bm 0}&s(t){\bm S}^\dagger&{\bm 0}
258: \end{array}\right]\,,
259: \end{equation}
260: for use with the Schr\"odinger equation
261: \begin{equation}
262: i\hbar\frac{d}{dt} {\bm C}(t)={\bm H}(t){\bm C}(t)\,.
263: \end{equation}
264: Here the zeros ${\bm 0}$ denote null square or rectangular matrices of
265: appropriate dimensions. The zero matrix in the bottom right corner
266: indicates that the system is supposed to maintain two-photon
267: resonance. All time dependence occurs in the two pulse amplitudes
268: $p(t)$ and $s(t)$, each with unit maximum value. The $N_e\times N_e$
269: diagonal matrix $\hbar\bm\Delta$ describes the detuning of the pump
270: carrier frequency from the Bohr frequency of the $g-e$ transition.
271: The $N_g\times N_e$ matrix $2 p(t){\bm P}/\hbar$ consists of Rabi
272: frequencies associated with the transitions between the $g$ and $e$
273: sets, $\hbar\Omega_{ij}(t)=2p(t)P_{ij}$. The elements of the constant
274: matrix ${\bm P}$ read
275: \begin{equation}
276: P_{ij}=\frac{1}{2}{\cal E}^{(p)}\mu_{ij}\,,\qquad
277: \left\{\begin{array}{l}
278: i=1\hdots N_g \\
279: j=1\hdots N_e
280: \end{array}\right.\,,
281: \end{equation}
282: where ${\cal E}^{(p)}$ is the peak amplitude of the pump-pulse
283: electric field and $\mu_{ij}$ is the dipole-transition moment.
284:
285: Similarly, the $N_e\times N_f$ matrix $2s(t){\bm S}/\hbar$ consists of
286: Rabi frequencies associated with the transitions between the $e$ and
287: $f$ sets of states. The elements of the constant matrix $\bm S$ are
288: \begin{equation}
289: S_{ij}=\frac{1}{2}{\cal E}^{(S)}\mu_{ij}\,,\qquad
290: \left\{\begin{array}{l}
291: i=1\hdots N_e \\
292: j=1\hdots N_f
293: \end{array}\right.\,,
294: \end{equation}
295: where ${\cal E}^{(S)}$ is the peak amplitude of the Stokes electric
296: field.
297:
298: The structure of the RWA Hamiltonian of Eq.~(\ref{ham}) is similar to
299: that of the conventional three-state STIRAP, in having all time
300: dependence confined to two pulses $p(t)$ and $s(t)$, but instead of
301: single ground, excited, and final states we have degenerate manifolds
302: of sublevels, and hence we have matrices $ p(t){\bm P}$, $s(t){\bm
303: S}$, and ${\bm \Delta}$ where conventional STIRAP would have scalar
304: elements. To illustrate these Fig. 1 shows the linkage patterns for
305: the angular momentum sequence $J = 2 \leftrightarrow 3 \leftrightarrow
306: 4$. To simplify the drawings we show the energies of successive
307: manifolds as increasing, such as would occur with a ladder scheme; the
308: connections are the same as with the usual lambda couplings, in which
309: the final sublevels have energies below the excited state.
310:
311: %---------------------------------------------------------
312: \begin{figure}
313: \includegraphics[width=\textwidth/2-1cm]{fig1} %fig1
314: \caption{(Color Online) An example for the degenerate STIRAP scheme:
315: we have three Zeeman multiplets with $J_g=2$, $J_e=3$, $J_f=4$.
316: The couplings are those of $\sigma_{\pm}$ polarized pulses. The
317: pump and Stokes pulses are detuned from exact resonance with the
318: excited-state by $\Delta$, but they maintain two-photon resonance
319: between states $g$ and $f$. The system separates into two
320: independent systems, indicated by solid and dashed lines. }
321: \label{fig:234-scheme}
322: \end{figure}
323: %---------------------------------------------------------
324:
325:
326: Although we discuss situations in which the coupling matrices result
327: from magnetic-sublevel degeneracy, all of our results apply quite
328: generally, for any mathematical form of the dipole-moment matrices and
329: consequently for any arbitrary structure of the constant matrices
330: ${\bm S}$ and ${\bm P}$.
331:
332: %------------------------------------------------------
333: \subsection{Dark states} \label{sec:dark-states}
334: %------------------------------------------------------
335:
336: There exist $N=N_g + N_e + N_f$ basis states for this system, and
337: hence $N$ adiabatic states $\advecb_n(t)$. We can immediately apply
338: the MS transformation \cite{Morris83}, at each instant of time, by
339: placing the $g$ and $f$ sets of states together into the MS $a$ set,
340: and taking the $e$ set to be the MS $b$ set. If the $a$ set is larger
341: than the $b$ set, there will be $N_u = N_a - N_b$ uncoupled states.
342: None of these have any component from the $e$ set, and so they are all
343: dark states. The number of dark states is thus $N_D = N_g + N_f -
344: N_e$. In the conventional nondegenerate STIRAP \cite{AAMOP}, for
345: which $N = 3$, the MS transformation gives one dark state and one
346: bright state; for the tripod system, for which $N = 4$, there are two
347: dark states \cite{Unanyan98,Theuer99}. In the angular-momentum system
348: of Fig. \ref{fig:234-scheme} there are $N_D = 5 + 9 - 7 = 7$ dark
349: states.
350:
351: For conventional nondegenerate STIRAP the composition of the dark
352: state changes with time, because the coupling matrices and the MS
353: transformation change with time. However, it is possible to associate
354: the (single) dark state initially with the nondegenerate ground state
355: by applying the pulses in the counterintuitive order, i.e. Stokes
356: pulse preceding pump pulse. When there is degeneracy, it is necessary
357: to establish that the entire population of any pure initial state in
358: the $g$ set is projected into the set of dark states and no population
359: is left in bright states. This completeness of the dark states is at
360: the heart of our question concerning the possibility of STIRAP with
361: degeneracy.
362:
363: %------------------------------------------------------
364: \section{General condition for complete population transfer}
365: \label{sec:mstrafo}
366: %------------------------------------------------------
367:
368: One of our basic questions is whether, for a given linkage pattern, it
369: is possible to empty completely the $g$ set for any arbitrary initial
370: state, once we have fixed the pump and Stokes pulses.
371:
372: It is easy to see that one necessary condition for complete removal of
373: population from the ground manifold is that there should not be more
374: sublevels in this manifold than there are in the excited state, i.e.
375: we require $N_g \leq N_e$
376:
377: To prove this assertion we employ a MS transformation \cite{Morris83}
378: on the pump transitions that connect ground and excited states. This
379: transformation introduces a new set of basis states in each of these
380: manifolds, such that each sublevel from the $g$ set couples to at most
381: one sublevel from the $e$ set. Were there are no Stokes couplings
382: between $e$ and $f$ states, the dynamics could be described as a set
383: of independent two-state systems, together with some single states
384: (uncoupled states) that are not affected by the pump radiation. Given
385: such a revision of the basis states, it is easy to see that if there
386: are more ground states than excited states, $N_g > N_e$, then the dark
387: states will be composed of $g$-states and some population will be
388: trapped there. This will remain unaffected by the radiation;
389: population cannot be removed from them using this particular linkage
390: pattern.
391:
392: Figure \ref{fig:212scheme} illustrates this accounting procedure. The
393: top frame (a) shows a general coupling scheme for the sequence $J= 2
394: \leftrightarrow 1 \leftrightarrow 2$. The MS transformation on the
395: $g-e$ pump transition produces the description shown in the bottom
396: frame (b). In the $g$ set, with this transformed basis, there occur
397: two sublevels that have no connection with any excited states.
398: Population cannot be removed from these as long as the couplings are
399: those shown in the top frame.
400:
401: It is easy to see that, had there been more sublevels in the $e$ set,
402: such that $N_g \le N_e$, then every one of the transformed states from
403: the $g$ set would be linked to some excited state, with consequent
404: possibility for population removal. There will also be uncoupled
405: states in the $e$ manifold but they are unpopulated and do not affect
406: the population transfer.
407:
408:
409: %---------------------------------------------------------
410: \begin{figure}
411: \includegraphics[width=5cm]{fig2} %fig2
412: \caption{(Color Online)
413: This sketch shows that when the number of ground-state-sublevels
414: exceeds the number of excited-state-sublevels then it is impossible
415: to transfer all the population from the ground-state manifold to the
416: final-state manifold. Frame~(a) shows the original coupling
417: configuration. Frame~(b) shows the couplings after a MS
418: transformation on the $g-e$ transition. The empty rectangles
419: represent uncoupled states. This transformation is independent of
420: time, because all elements of the coupling share a common time
421: dependence, $p(t)$. The presence of uncoupled sublevels in the $g$
422: set prevents removal of population from these states; hence it is
423: not possible to remove all population from all of the ground-state
424: sublevels. }
425: \label{fig:212scheme}
426: \end{figure}
427: %---------------------------------------------------------
428:
429: The introduction of MS basis states in this way makes the $g-e$
430: linkage pattern quite simple, but by introducing a new basis the $e-f$
431: couplings become more complicated: generally there will be a
432: connection between each transformed $e$ state and each (untransformed)
433: $f$ state, as indicated in frame (b).
434:
435: Next we consider the $e-f$ coupling. We can repeat the previous
436: argument for the $g-e$ coupling with the replacements
437: $g\leftrightarrow e$ and $e\leftrightarrow f$. We obtain, that the
438: Stokes field MS transformation yields $N_{<}=\mbox{min}\{N_e, N_f\}$
439: independent two level systems for the $e-f$ transition, plus
440: $|N_e-N_f|$ uncoupled states in the larger one out of the $e$ and $f$
441: sets. It is easy to see that, had there been more sublevels in the $f$
442: set, such that $N_e \le N_f$, then every one of the transformed states
443: from the $e$ set would be linked to some final state, with consequent
444: possibility for population removal. Combinig the arguments of the MS
445: tarnsormations for the $g-e$ and $e-f$ couplings, we obtain that in
446: general, if a non--descending sequence of state--degeneracies is
447: fulfilled
448: \begin{equation}\label{cond}
449: N_g\leq N_e \leq N_f\,,
450: \end{equation}
451: then a complete STIRAP-like population transfer from the $g$ set to
452: the $f$ set is possible. We emphasize that in this case the success
453: of the full transfer is independent of the initial state of the
454: system: it can be any pure state or a mixed state as well.
455:
456:
457: A particularly important special case of degeneracy occurs when there
458: are dark states but they are insufficient to produce complete
459: population transfer. This occurs when $N_g+N_f>N_e$, but $N_f<N_e$.
460: For example, in the linkage of $J=1\leftrightarrow 2 \leftrightarrow
461: 1$ there is 1 dark state. Figure~\ref{fig:121scheme} illustrates this
462: situation.
463:
464: %------------------------------------------------------
465: \section{The Stokes-field MS transformation} \label{sec:stokes-mstrafo}
466: %------------------------------------------------------
467:
468: In this section we determine the dark states of the Hamiltonian of
469: Eq.~(\ref{ham}); these are the adiabatic states that will be utilized
470: for the desired adiabatic population transfer. In order to simplify
471: the structure of the Hamiltonian, we perform a MS transformation; here
472: we take that to be on the $e-f$ couplings (those of the Stokes field).
473: In our case the time-independent transformation matrix ${\bm U}$ is
474: defined as
475: \begin{equation}\label{Udef}
476: {\bm U}=\left[
477: \begin{array}{ccc}
478: {\bm I}&{\bm 0}&{\bm 0}\\
479: {\bm 0}&{\bm B}&{\bm 0}\\
480: {\bm 0}&{\bm 0}&{\bm A}
481: \end{array}\right]\,.
482: \end{equation}
483: In the top-left corner there is a unit matrix ${\bm I}$ of dimension
484: $N_g\times N_g$. This leaves the $g$ set of states unaltered. The
485: $N_f\times N_f$ unitary matrix $\bm A$ transforms the sublevels in the
486: final-state manifold. Similarly, the $N_e\times N_e$ unitary matrix
487: $\bm B$ transforms the sublevels in the excited-state manifold. The
488: constant matrices $\bm A$ and $\bm B$ are defined \cite{Morris83} such
489: that by transforming the Hamiltonian Eq.~(\ref{ham}) with the matrix
490: $\bm U$ through the relation
491: \begin{equation}\label{trafo}
492: {\bm U}{\bm H}(t){\bm U}^\dagger = \left[\begin{array}{ccc}
493: {\bm 0}&p(t)\widetilde{\bm P}&{\bm 0}\\
494: p(t)\widetilde{\bm P}^\dagger&\hbar{\bm\Delta}
495: &s(t)\widetilde{\bm S}\\
496: {\bm 0}&s(t)\widetilde{\bm S}^\dagger
497: \end{array}\right]\,
498: \end{equation}
499: we obtain a transformed pump-field coupling matrix $\widetilde{\bm
500: P}={\bm P}{\bm B}^\dagger$, and a quasi-diagonal Stokes-field
501: coupling matrix $\widetilde{\bm S}=\bm B\bm S\bm A^\dagger$. By
502: quasi-diagonal we mean that the structure of the matrix is
503: \begin{eqnarray}\label{Str}
504: \widetilde{\bm S} =
505: \begin{cases} \left[\begin{array}{cc}
506: \widetilde{\bm \Sigma} &
507: {\bm 0} \end{array}\right] &
508: \mbox{ if }N_f>N_e\,,
509: \\[6pt]
510: \widetilde{\bm \Sigma} &
511: \,\,\mbox{if } N_f=N_e\,,
512: \\[2pt]
513: \left[\begin{array}{c}
514: \widetilde{\bm \Sigma} \\ {\bm 0}
515: \end{array}\right] &
516: \mbox{ if } N_f<N_e\,,
517: \end{cases}
518: \end{eqnarray}
519: where $\widetilde{\bm \Sigma}$ is a square diagonal matrix with
520: dimension $N_<={\rm min}(N_e, N_f)$. The moduli of the diagonal
521: elements are given by the square-roots of the common eigenvalues of
522: the Hermitian matrices $\bm S\bm S^{\dag}$ (of dimension $N_e\times
523: N_e$) and $\bm S^{\dag}\bm S$ (of dimension $N_f\times N_f$). The
524: phases of the diagonal elements are obtained by evaluating directly
525: the matrix product $\bm B\bm S\bm A^\dagger$. Some of the diagonal
526: elements of $\widetilde{\bm \Sigma}$ might be zero, meaning that some
527: $e-f$ couplings vanish in the MS basis. We here assume that in
528: general all diagonal elements of $\widetilde{\bm\Sigma}$ are non-zero,
529: i.e. it is nonsingular. We treat in Appendix \ref{sec:sing-sigma} the
530: case when this matrix is singular.
531:
532:
533: In the following subsections we consider the three important special
534: cases of degeneracies and derive the adiabatic states of the coupled
535: degenerate systems.
536:
537:
538: %------------------------------------------------------
539: \subsection{The case $ N_g\leq N_e\leq N_f$}\label{sec:piramid1}
540: %------------------------------------------------------
541:
542: %---------------------------------------------------------
543: \begin{figure}
544: \includegraphics[width=5cm]{fig3} %fig3
545: \caption{(Color Online)
546: The three stages of the transformations. (a) The original
547: coupling scheme. (b) The result of the Stokes-field MS
548: transformation, converting the couplings between $e$ and $f$
549: sets into independent one- and two-state systems. (c) The
550: result of redefining the states in the $g$, $e$, and $f$ sets.
551: }
552: \label{fig:112scheme}
553: \end{figure}
554: %---------------------------------------------------------
555:
556:
557: We first consider the case when the MS transformation on the $e-f$
558: transition results in $N_f-N_e>0$ decoupled sublevels in the $f$
559: manifold. The coupling matrix $\widetilde{\bm S}$ takes the form
560: given in the first row of Eq.~(\ref{Str}), and hence the Hamiltonian
561: in the MS basis reads
562: \begin{equation}\label{ham-ms}
563: \widetilde{{\bm H}}(t)=\left[\begin{array}{cccc}
564: {\bm 0}&p(t)\widetilde{{\bm P}}&{\bm 0}&{\bm 0}\\
565: p(t)\widetilde{{\bm P}}^\dagger&\hbar{\bm
566: \Delta}&s(t)\widetilde{\bm \Sigma}
567: &{\bm 0}\\
568: {\bm 0}&s(t)\widetilde{\bm \Sigma}^\dagger&{\bm 0}&{\bm 0}\\
569: {\bm 0}&{\bm 0}&{\bm 0}&{\bm 0}
570: \end{array}\right]\,.
571: \end{equation}
572: As with the original RWA Hamiltonian, the only time dependence enters
573: through the pulses $p(t)$ and $s(t)$.
574:
575: We can treat the system in the same way when $N_f=N_e$. Then the
576: coupling matrix $\widetilde{\bm S}$ is given by the second row of
577: Eq.~(\ref{Str}), and we have to omit all zero rows and columns from
578: the Hamiltonian of Eq.~(\ref{ham-ms}). In either cases the sub-matrix
579: $\widetilde{{\bm P}}$ has dimensions $N_g\times N_e$, while the square
580: matrices $\widetilde{\bm \Sigma}$ and $\bm\Delta$ have dimensions
581: $N_e\times N_e$.
582:
583: To find the adiabatic eigenvectors $\widetilde{\advecb}_k(t)$ of
584: $\widetilde{{\bm H}}(t)$ we take their elements to have the form
585: \begin{equation}
586: \label{vkparam}
587: \widetilde{\advecb}_k(t) = \left[\begin{array}{c}
588: {\bm x}_k \t \\
589: \widetilde{\bm y}_k \t \\
590: \widetilde{\bm z}_k \t \\
591: \widetilde{\bm z}'_k \t
592: \end{array}\right]\quad
593: \begin{array}{c}
594: {g} \\ {e} \\ {f} \\ {f'}
595: \end{array}
596: \end{equation}
597: where $f'$ denotes the subspace of uncoupled states in the $f$ set.
598: Because these are unlinked to the $e$ set they meet the definition of
599: dark states. Their population, if initially present, is preserved
600: throughout the time evolution. When $N_f=N_e$ we simply omit the
601: fourth row from this vector (the $f'$ states), i.e. we do not have
602: $\widetilde{\bm z}'_k$. In Eq.~(\ref{vkparam}) there is no tilde on
603: the $x$ components because, unlike the $y$ and $z$ components, these
604: do not transform in the Stokes field MS transformation. In Sec. IV B
605: and C the $x$ components undergo a MS transformation, as is indicated
606: there by a tilde.
607:
608: The eigenvectors satisfy the eigenvalue equation
609: \begin{equation}\label{eigeneqdef}
610: \widetilde{\bm {H}}(t)\widetilde{\advecb}_k(t)=\eigenv_k(t)
611: \widetilde{\advecb}_k(t)\,.
612: \end{equation}
613: By substituting the Hamiltonian of Eq.~(\ref{ham-ms}) and the
614: parameterization (\ref{vkparam}) of the eigenvectors into this
615: equation we obtain four sets of coupled linear equations for ${\bm
616: x}_k$, $\widetilde{\bm y}_k$, $\widetilde{\bm z}_k$, and
617: $\widetilde{\bm z}'_k$. The solution of these equations provide the
618: dark and bright eigenvectors $\widetilde{\advecb}_k(t)$ defined by
619: Eq.~(\ref{eigeneqdef}).
620:
621: Let us assume that there exists an eigenvalue zero, $\eigenv_0=0$\,.
622: This is always possible to ensure, by suitable choice of the phases of
623: the rotating wave approximation and the zero-point of energy. If we
624: can find a solution of the eigenvalue-equation (\ref{eigeneqdef}) for
625: this case, then our assumption $\eigenv_0=0$ holds, since the solution
626: of the linear equations is unique. After some algebra one can obtain
627: $N_g$ different vectors $\widetilde{\advecb}^{(l)}_0 \t\,\,,l=1\ldots
628: N_g$, that are linearly independent of each other, and can make these
629: orthonormal
630: \begin{equation}\label{darkstates1}
631: \widetilde{\advecb}_0^{(l)}(t)= {1\over {\cal N}_0^{(l)}(t)}\left[
632: \begin{array}{c}
633: s(t){\bm x}_0^{(l)} \\ {\bm 0} \\
634: -p(t)\widetilde{\bm \Sigma}^{-1}\widetilde{{\bm
635: P}}^\dagger{\bm x}_0^{(l)} \\
636: {\bm 0}
637: \end{array}\right]\,,
638: \end{equation}
639: where ${\cal N}_0^{(l)} (t)$ is a (time dependent) normalization
640: factor. Here we have assumed that the matrix $\widetilde{\bm \Sigma}$
641: is nonsingular. We will discuss separately, in Appendix
642: \ref{sec:sing-sigma}, the situation when $\widetilde{\bm \Sigma}$ is
643: singular. Since the $y$ component of these vectors is zero, they
644: have no component in the $e$ set; they correspond to dark states. To
645: make the dark eigenvectors of Eq.~(\ref{darkstates1}) orthogonal we
646: require that
647: \begin{equation}\label{ortho}
648: s(t)^2 \langle{\bm x}^{(k)\,T}_0|{\bm x}^{(l)}_0\rangle+
649: p(t)^2\langle{\bm x}^{(k)\,T}_0|
650: \widetilde{{\bm P}}\widetilde{\bm \Sigma}^{-1\dagger}
651: \widetilde{\bm \Sigma}^{-1}\widetilde{{\bm P}}^\dagger
652: |{\bm x}^{(l)}_0\rangle=0\,,
653: \end{equation}
654: for $1 \leq k < l \leq N_g$. The time-dependence of the envelope
655: functions $s(t)$ and $p(t)$ is arbitrary, and therefore we require
656: that the two terms on the left-hand-side (lhs) of Eq.~(\ref{ortho}) be
657: identically zero. The eigenvectors of a Hermitian matrix can be
658: chosen so that they are orthogonal to each-other, and therefore the
659: first term on the lhs of Eq.~(\ref{ortho}) is automatically zero. It
660: follows that the vectors ${\bm x}^{(l)}_0$ are the eigenvectors of the
661: Hermitian matrix
662: \begin{equation}\label{metric}
663: {\bm M} = {\bm P}({\bm S}{\bm S}^{\dagger})^{-1}{\bm P}^{\dagger}\equiv
664: \widetilde{{\bm P}}\widetilde{\bm
665: \Sigma}^{-1\dagger}\widetilde{\bm \Sigma}^{-1}
666: \widetilde{{\bm P}}^\dagger\,.
667: \end{equation}
668:
669: There is another set of dark eigenvectors for $N_f>N_e$. These follow
670: from the discussion after Eq.~(\ref{vkparam}) and are given by
671: \begin{equation}
672: \widetilde{\advecb}_0^{(l)}=\left[
673: \begin{array}{c}
674: {\bm 0} \\ {\bm 0} \\ {\bm 0} \\ {\bm z}^{\prime\, (l)}
675: \end{array}\right]\,,\quad l=N_g+1, \ldots N_f-N_e+N_g\,,
676: \end{equation}
677: where ${\bm z}^{\prime\, (l)}$ are constant orthonormal unit vectors.
678: These dark eigenvectors are clearly orthogonal to those of
679: Eq.~(\ref{darkstates1}).
680:
681: We show in Appendix~\ref{sec:linearized-couplings} that the coupling
682: sequence $g\leftrightarrow e\leftrightarrow f$ can be rendered to
683: independent three-state chains by a suitable hoice of the basis states
684: in the $g$, $e$, and $f$ sets. Figure~\ref{fig:112scheme} illustrates
685: the sequence of transformations that leads to the construction of the
686: dark-state eigenvectors Eq.~(\ref{darkstates1}). Frame (a) shows the
687: original system, with some couplings. Frame (b) depicts the results
688: of the Stokes-field MS transformation of the $e$ and $f$ states.
689: Frame (c) shows the result of the redefinition of the $g$, $e$, and
690: $f$ sets of states according to Eq.~(\ref{sets}), with the resulting
691: set of independent chains.
692:
693:
694: The matrix of Eq.~(\ref{metric}) may have zero eigenvalues as well.
695: If so, the corresponding eigenvectors ${\bm x}_0^{(k)}$ satisfy the
696: equation
697: \begin{equation}\label{coupcond}
698: {\bm P}^{\dagger}{\bm x}_0^{(k)}={\bm 0}\,,
699: \end{equation}
700: since we have assumed that the matrix $\widetilde{\bm \Sigma}$ is
701: nonsingular. Note that here ${\bm P}^{\dag}$ is expressed in the
702: original atomic basis. The $i$th row of the matrix ${\bm
703: P}^{\dagger}$ describes the coupling between state $i$ from the $e$
704: set and the sublevels of the $g$ set. The rows of the coupling matrix
705: can be considered as vectors that span a subspace of states from the
706: $g$ set. The dimension of this subspace is the number of linearly
707: independent rows of ${\bm P}^{\dagger}$, say $N_P$. Obviously we have
708: $N_P\leq N_e$ and $N_P\leq N_g$. Therefore, there are $N_g-N_P$
709: different, nontrivial solutions of Eq.~(\ref{coupcond}). These
710: nontrivial solutions provide states that are unaffected by the pump
711: field.
712:
713: If $N_g=N_P\leq N_e$ then such an uncoupled state does not exist, and
714: the vectors $\{{\bm x}_0^{(k)}\}$, $k=1\ldots N_g$ span the total
715: $g$-set manifold. Therefore by choosing a counterintuitive
716: pulse-sequence for the pump and Stokes pulses, we can cause complete
717: transfer of population from the $g$ set to the $f$ set by means of
718: independent STIRAP processes. For such population transfer to
719: succeed, the conditions of the adiabatic evolution should be
720: fulfilled, as we will discuss in Sec.~\ref{sec:adiab-tevol}. The
721: success of such population transfer is independent of the initial
722: state of the system. It can be any single state, an arbitrary
723: coherent superposition of states or even a mixed state, {see
724: Sec.~\ref{sec:adiab-tevol}}.
725:
726: If $N_P<N_g$ then some $g$-set sublevels are decoupled from the pump
727: field, hence in general it is then impossible to move all the
728: population from the $g$ set. Part of it is trapped in dark states.
729:
730: The other $2N_e$ adiabatic eigenvectors belong to non-zero
731: eigenvalues. They can be obtained in the form
732: \begin{equation}
733: \widetilde{\advecb}_k \t= {1\over {\cal N}_k(t)}\left[\begin{array}{c}
734: p(t)\widetilde{{\bm P}} \widetilde{\bm y}_k \t\\
735: \eigenv_k \t \widetilde{\bm y}_k \t\\
736: s(t)\widetilde{\bm \Sigma}^{\dagger} \widetilde{\bm y}_k \t\\
737: {\bm 0}
738: \end{array}\right]\,,\qquad k=1\ldots 2N_e
739: \label{eq:vk}
740: \end{equation}
741: where ${\cal N}_k(t)$ is a normalization factor and $\widetilde{\bm y}_k
742: \t$ satisfies the eigenvalue equation
743: \begin{equation}
744: \left[ p(t)^2\widetilde{{\bm P}}^\dagger \widetilde{{\bm P}} + v(t)^2
745: \widetilde{\bm \Sigma}\widetilde{\bm \Sigma}^{\dagger}\right]
746: \widetilde{\bm y}_k \t =
747: \eigenv_k\t [\eigenv_k\t-\hbar\Delta] \widetilde{\bm y}_k \t\,.
748: \label{eq:eigen}
749: \end{equation}
750: Because they contain component states from the $e$ set, these are
751: bright states. Although for population transfer we use the dark
752: states of Eq.~(\ref{darkstates1}), we need the bright states to find
753: the adiabaticity conditions; see Sec.~\ref{sec:adiab-tevol}.
754:
755:
756: In summary: in this subsection we have shown that when $N_g\leq
757: N_e\leq N_f$, under very general conditions the complete population
758: from the $g$ set can be transferred to the $f$ set of states. Once we
759: have fixed the pulse-shapes, polarizations and phases, complete
760: transfer can be obtained for any arbitrary initial state from the $g$
761: set. The eigenvectors $\advecb_k\t$, $k=0\ldots 2N_e$ in the original
762: bare atomic basis can be obtained as
763: \begin{equation}\label{advecb}
764: \advecb_k\t = \frac{1}{{\cal N}_k\t}{\bm U}^\dagger\left[ \begin{array}{c}
765: {\bm x}_k \t \\
766: \widetilde{\bm y}_k\t \\
767: \widetilde{\bm z}_k\t \\
768: {\bm 0} \\
769: \end{array} \right] =\frac{1}{{\cal N}_k\t}
770: \left[ \begin{array}{c}
771: {\bm x}_k \t\\
772: {\bm B}^\dagger \widetilde{\bm y}_k\t\\
773: {\bm A}^\dagger \left[ \begin{array}{c}
774: \widetilde{\bm z}_k\t \\
775: {\bm 0}
776: \end{array} \right]
777: \end{array} \right]\,.
778: \end{equation}
779: Moreover, with this method it is possible not only to transfer
780: populations, but to create superposition states in the $f$ set. We
781: will consider this possibility in Sec.~\ref{sec:examples}.
782:
783: %------------------------------------------------------
784: \subsection{The case $N_g>N_e>N_f$}\label{sec:piramid2}
785: %------------------------------------------------------
786:
787:
788: According to the considerations presented in the beginning of
789: Sec.~\ref{sec:mstrafo}, we cannot expect that all the population from
790: the $g$ set can be removed when $N_g>N_e>N_f$. However, a part of the
791: population can be removed and with this we can create coherent
792: superposition states in the $f$ set. In order to find the dark- and
793: bright states of the system we proceed in the same way as in
794: Sec.~\ref{sec:piramid1}, but now with the MS transformation involving
795: the pump transition
796: \begin{equation}\label{Udef2}
797: {\bm U}=\left[
798: \begin{array}{ccc}
799: {\bm B}&{\bm 0}&{\bm 0}\\
800: {\bm 0}&{\bm A}&{\bm 0}\\
801: {\bm 0}&{\bm 0}&{\bm I}
802: \end{array}\right]\,.
803: \end{equation}
804: We look for the eigenvectors of the transformed Hamiltonian in the form
805: \begin{equation}
806: \label{vkparam2}
807: \widetilde{\advecb}_k \t = \left[\begin{array}{c}
808: \widetilde{\bm x}_k \t \\
809: \widetilde{\bm x}^{\prime}_k \t \\
810: \widetilde{\bm y}_k \t \\
811: {\bm z}_k \t
812: \end{array}\right]\,.\quad
813: \begin{array}{c}
814: {g} \\ {g'} \\ {e} \\ {f}
815: \end{array}
816: \end{equation}
817: The vectors $\widetilde{\bm x}^{\prime}_k \t$ describe the population
818: in those states of the $g$ set that are decoupled from the pump field.
819: There are $N_g-N_e$ dark states in the $g$ manifold, and these can be
820: written in the form
821: \begin{equation}
822: \label{vkparam2b}
823: \widetilde{\advecb}_0^{(l)} \t= \left[\begin{array}{c}
824: {\bm 0}\\
825: \widetilde{\bm x}^{\prime (l)}_0 \t \\
826: {\bm 0}\\
827: {\bm 0}
828: \end{array}\right]\,,
829: \end{equation}
830: where the vectors $\{\widetilde{\bm x}^{\prime (l)}_0 \t\} $ form
831: an orthonormal set.
832: The population cannot be removed from these states. The rest of the
833: dark states are obtained in the manner used for
834: Eq.~(\ref{darkstates1}). They can be written as
835: \begin{equation}
836: \widetilde{\advecb}^{(k)}_0 \t = {1\over {\cal N}^{(k)}_0 \t} \left[
837: \begin{array}{c}
838: s(t)\widetilde{\bm \Pi}^{\dag -1}\widetilde{{\bm S}}{\bm z}^{(k)}_0 \\
839: {\bm 0} \\
840: {\bm 0} \\
841: -p(t){\bm z}^{(k)}_0
842: \end{array}\right]\,,
843: \label{darkstates2}
844: \end{equation}
845: where $\left[\begin{array}{c} \widetilde{\bm \Pi}\\ {\bm 0}
846: \end{array} \right] =\bm B\bm P\bm A^\dagger$, with $\widetilde{\bm
847: \Pi}$ a diagonal coupling matrix of dimension $N_e\times N_e$, and
848: $\widetilde{{\bm S}}={\bm A}{\bm S}$. We require orthogonality for
849: the dark states Eq.~(\ref{darkstates2}). Hence the constant vectors
850: ${\bm z}^{(k)}_0$ are chosen so that they are eigenstates of the
851: Hermitian matrix $\widetilde{{\bm S}}^{\dagger}\widetilde{\bm
852: \Pi}^{-1} \widetilde{\bm \Pi}^{\dagger -1}\widetilde{{\bm S}}$, in
853: direct analogy with the way the constant vectors ${\bm x}^{(k)}_0$
854: were chosen earlier in Sec.~\ref{sec:piramid1}.
855:
856:
857:
858:
859: %------------------------------------------------------
860: \subsection{The case $N_g,N_f<N_e$}\label{sec:diamond}
861: %------------------------------------------------------
862:
863: Here we consider the situation $N_g,N_f<N_e$. We will show that under
864: these conditions the dark states of the system can be identified by
865: means of two sequential MS transformations. The first MS
866: transformation is performed among the $e$ and $f$ sets of the Stokes
867: transition, as in subsection \ref{sec:piramid1}. The transformation
868: matrix is given by Eq.~(\ref{Udef}). As a result, the coupling matrix
869: ${\bm S}$ of the Hamiltonian (\ref{ham}) takes the quasi-diagonal form
870: of the third row of Eq.~(\ref{Str}). Therefore, the Hamiltonian in
871: the MS basis reads
872: \begin{equation}\label{Hamb}
873: \widetilde{{\bm H}}(t)=\left[\begin{array}{cccc}
874: {\bm 0}&p(t)\widetilde{{\bm P}}_a&p(t)\widetilde{{\bm P}}_b&{\bm 0}\\
875: p(t)\widetilde{{\bm P}}_a^\dagger&\hbar{\bm \Delta}&{\bm 0}&
876: s(t)\widetilde{\bm \Sigma}\\
877: p(t)\widetilde{{\bm P}}_b^{\prime\dagger}&{\bm 0}&\hbar{\bm
878: \Delta}&{\bm 0}\\
879: {\bm 0}&s(t)\widetilde{\bm \Sigma}^\dagger&{\bm 0}&{\bm 0}\\
880: \end{array}\right]\,.
881: \end{equation}
882: The diagonal square matrix $\widetilde{\bm \Sigma}$ has dimension
883: $N_f\times N_f$. It can be readily seen that there are $N_e-N_f$
884: states in the $e$ set that are not coupled to the $f$ set. We call
885: these {\em uncoupled} levels, whereas the other subset of coupled
886: excited-state-sublevels are called {\em active}. In the Hamiltonian
887: of Eq.~(\ref{Hamb}) the pump coupling matrix is partitioned into two
888: sub-matrices: the matrix $\widetilde{{\bm P}}_a$ of dimension
889: $N_g\times N_f$ describes couplings between the $g$ set and the active
890: MS states of the $e$ set. The other sub-matrix $\widetilde{{\bm
891: P}}_b$ of dimension $N_g\times (N_e-N_f)$ is associated with the
892: transitions between the states of the $g$ set and the uncoupled states
893: of the $e$ set. The result of this transformation is illustrated in
894: Fig.~(\ref{fig:212b}b). As the figure shows, we cannot identify
895: clearly the dark states, because in general all the states of the $g$
896: set are coupled to all of the $e$ set.
897: %---------------------------------------------------------
898: \begin{figure}
899: \includegraphics[width=5cm]{fig4} %fig4
900: \caption{(Color Online) The three frames show the stages of the
901: MS transformations when the sequence of degenerate states violates
902: the condition Eq.~(\ref{cond}): Frame (a) depicts the original
903: coupling scheme , here $N_f<N_e$. Frame (b) shows the result of the
904: Stokes field MS transformation, converting the couplings between $e$
905: and $f$ sets into independent one- and two-state systems. The
906: one-state systems are in the $e$ set. Frame (c) shows the result of
907: the pump field MS transformation followed by the redefinition of the
908: sublevels in the $g$, $e$, and $f$ sets according to
909: Eq.~(\ref{sets}), leading to one dark state. This dark state is
910: associated with the middle three-state linkage, indicated by heavy
911: lines. In addition, the middle $e$ state is coupled not only to a
912: single $g$ state but to the two others as well. As a result, the
913: populations in the two spectator $g$ states may disturb the complete
914: population transfer from the middle $g$ state, see text. }
915: \label{fig:212b}
916: \end{figure}
917: %---------------------------------------------------------
918: Therefore, we perform a second MS transformation, involving the $g$
919: set and just those states of the $e$ set that are decoupled from the
920: $f$ set -- two in the present example. The result is illustrated in
921: Fig.~(\ref{fig:212b}c). In this example there is one $g$-set state
922: that couples solely to an active MS state of the $e$ set because the
923: other two have, by means of the MS transformation, been linked to the
924: two uncoupled $e$ states. The population can be moved from this $g$
925: state to an $f$ state. The middle $e$ state is coupled to all three
926: $g$ states. Consequently, if the two spectator $g$ states are
927: populated, they disturb the complete population transfer from the
928: middle $g$ state into $f$ states, and the population transfer process
929: will place population into the $e$ and $f$ states.
930:
931:
932: In general, the transformation matrix of the second MS transformation
933: is defined as
934: \begin{equation}\label{Updef}
935: {\bm U}'=\left[
936: \begin{array}{cccc}
937: {\bm A}'&{\bm 0}&{\bm 0}&{\bm 0}\\
938: {\bm 0}&{\bm I}&{\bm 0}&{\bm 0}\\
939: {\bm 0}&{\bm 0}&{\bm B}'&{\bm 0}\\
940: {\bm 0}&{\bm 0}&{\bm 0}&{\bm I}
941: \end{array}\right]\,.
942: \end{equation}
943: The $N_g\times N_g$ unitary matrix ${\bm A}'$ transforms the $g$ set,
944: whereas the $(N_e-N_f)\times (N_e-N_f)$ unitary matrix ${\bm B}'$
945: transforms the uncoupled states of the $e$ set. The two unit matrices
946: ${\bm I}$ are of dimension $N_f\times N_f$. For $N_e-N_f<N_g$ the
947: transformation yields
948: \begin{subequations}\label{pi1}
949: \begin{eqnarray}
950: {\bm A}'\widetilde{\bm P}_a &=&
951: \left[\begin{array}{c}
952: \widetilde{\bm P} \\ \widetilde{\bm P}'
953: \end{array}\right]\,, \\
954: {\bm A}'\widetilde{\bm P}_b {\bm B}^{\prime\dag} &=&
955: \left[\begin{array}{c}
956: {\bm 0} \\ \widetilde{\bm \Pi}
957: \end{array}\right]\,,
958: \end{eqnarray}
959: \end{subequations}
960: where the matrix $\widetilde{\bm P}$ is of dimension
961: $[N_g-(N_e-N_f)]\times N_f$, $\widetilde{\bm P}'$ is of dimension
962: $(N_e-N_f)\times N_f$, and $\widetilde{\bm \Pi}$ is a diagonal matrix
963: of dimension $(N_e-N_f)\times (N_e-N_f)$. For $N_e-N_f=N_g$ we find
964: \begin{subequations}\label{pi2}
965: \begin{eqnarray}
966: {\bm A}'\widetilde{\bm P}_a &=& \widetilde{\bm P}'\,, \\
967: {\bm A}'\widetilde{\bm P}_b {\bm B}^{\prime\dag} &=&
968: \widetilde{\bm \Pi}\,,
969: \end{eqnarray}
970: \end{subequations}
971: where the matrix $\widetilde{\bm P}'$ is of dimension $N_g\times N_f$
972: and $\widetilde{\bm \Pi}$ is of dimension $N_g\times N_g$. We do not
973: have $\widetilde{\bm P}$ in this case. Finally, for $N_e-N_f>N_g$ we
974: get
975: \begin{subequations}\label{pi3}
976: \begin{eqnarray}
977: {\bm A}'\widetilde{\bm P}_a &=& \widetilde{\bm P}'\,, \\
978: {\bm A}'\widetilde{\bm P}_b {\bm B}^{\prime\dag} &=&
979: \left[\begin{array}{cc}
980: \widetilde{\bm \Pi}&{\bm 0}
981: \end{array}\right]\,,
982: \end{eqnarray}
983: \end{subequations}
984: where the matrix $\widetilde{\bm P}'$ is of dimension $N_g\times N_f$
985: and $\widetilde{\bm \Pi}$ is of dimension $N_g\times N_g$. Just as
986: with the conditions $N_e-N_f=N_g$, we do not have a matrix
987: $\widetilde{\bm P}$ in the present case either.
988:
989:
990: In general, none of the diagonal elements of the matrix
991: $\widetilde{\bm \Pi}$ are zero. Therefore, when $N_e-N_f\geq N_g$
992: there are no states in the $g$ set that are coupled solely to active
993: MS states in the $e$ set. It follows that no dark state can be
994: identified in the system and hence a STIRAP-like population transfer
995: is impossible.
996:
997: In special (but important) cases it may occur that some diagonal
998: elements of $\widetilde{\bm \Pi}$ vanish. Then the system has such MS
999: $g$-set states that are coupled only to active MS states of the $e$
1000: set, hence the system has dark states and a STIRAP process is
1001: possible. We will reconsider this case later in this subsection.
1002:
1003: In all three cases $N_e-N_f>N_g$, $N_e-N_f=N_g$, and $N_e-N_f<N_g$
1004: some initial population of the $g$ set cannot be included in the dark
1005: states of the system in the general case of arbitrary initial
1006: superposition of $g$ states. We conclude that in general in the case
1007: of $N_g, N_f<N_e$ it is impossible to remove all the population from
1008: the $g$ set in a STIRAP-like population transfer process. Exceptions
1009: occur when the matrix $\widetilde{\bm\Pi}$ is identically zero. Then
1010: the uncoupled MS states of the $e$ set are decoupled not only from the
1011: $f$ set, but also from the $g$ set.
1012:
1013: When $N_e-N_f<N_g$ the second MS transformation produces from the
1014: Hamiltonian of Eq.~(\ref{Hamb}) the matrix
1015: \begin{equation}\label{Hambt}
1016: {\widetilde{{\bm H}}}'(t)=\left[\begin{array}{ccccc}
1017: {\bm 0}&{\bm 0}&p(t)\widetilde{{\bm P}}&{\bm 0}&{\bm 0}\\
1018: {\bm 0}&{\bm 0}&p(t)\widetilde{{\bm P}}'&p(t)\widetilde{{\bm\Pi}} &
1019: {\bm 0}\\
1020: p(t)\widetilde{{\bm P}}^\dagger&p(t)\widetilde{{\bm P}}^{\prime\dagger}
1021: &\hbar{\bm \Delta}&{\bm 0}&s(t)\widetilde{\bm \Sigma}\\
1022: {\bm 0}&p(t)\widetilde{{\bm\Pi}}^{\dag}&{\bm 0}&\hbar{\bm \Delta}&{\bm
1023: 0}\\
1024: {\bm 0}&{\bm 0}&s(t)\widetilde{\bm \Sigma}^\dagger&{\bm
1025: 0}&{\bm 0}\\
1026: \end{array}\right]\,.
1027: \end{equation}
1028: The situation $N_e-N_f\geq N_g$ can be treated similarly. In order to
1029: find the dark states of the Hamiltonian of Eq.~(\ref{Hambt}) we
1030: proceed in the same way as in Sec.~\ref{sec:piramid1}. The
1031: eigenvectors are parameterized as
1032: \begin{equation}
1033: \label{vkparam2c}
1034: \widetilde{\advecb}_k \t=
1035: \left[\begin{array}{c}
1036: \widetilde{\bm x}_k \t \\
1037: \widetilde{\bm x}'_k \t \\
1038: \widetilde{\bm y}_k \t \\
1039: \widetilde{\bm y}'_k \t \\
1040: \widetilde{\bm z}_k \t\\
1041: \end{array}\right]\,.\quad
1042: \begin{array}{c}
1043: {g} \\ {g'} \\ {e} \\ {e'} \\ {f}
1044: \end{array}
1045: \end{equation}
1046: The eigenvalue equation is defined by Eq.~(\ref{eigeneqdef}). By
1047: inserting the Hamiltonian of (\ref{Hambt}) and the parameterization of
1048: the eigenvector Eq.~(\ref{vkparam2c}) into Eq.~(\ref{eigeneqdef}) we
1049: obtain five groups of coupled linear equations for $\widetilde{\bm
1050: x}_k$, $\widetilde{\bm x}'_k$, $\widetilde{\bm y}_k$,
1051: $\widetilde{\bm y}'_k$, and $\widetilde{\bm z}_k$. The $N_D$ linearly
1052: independent dark eigenvectors $\widetilde{\advecb}_0^{(l)}(t)$ that
1053: belong to the eigenvalue $\eigenv_0=0$, are given by
1054: \begin{equation}
1055: \widetilde{\advecb}_0^{(l)}(t) = {1\over {\cal N}_0 \t}\left[
1056: \begin{array}{c}
1057: s(t)\widetilde{\bm x}_0 \\ s(t)\widetilde{\bm x}'_0 \\ {\bm
1058: 0} \\ {\bm 0} \\
1059: -p(t)\widetilde{\bm \Sigma}^{-1}[\widetilde{{\bm
1060: P}}^\dagger\widetilde{\bm x}_0 +
1061: \widetilde{{\bm P}}^{\prime\dagger}\widetilde{\bm x}'_0]
1062: \end{array}\right]\,,
1063: \label{darkstates3}
1064: \end{equation}
1065: where, the constant vector $\widetilde{\bm x}'_0$ should satisfy the
1066: extra condition
1067: \begin{equation}\label{zero-subspace}
1068: \widetilde{\bm \Pi}\widetilde{\bm x}'_0={\bm 0}\,.
1069: \end{equation}
1070: This condition says that in a dark state no population can be in those
1071: $g$-set states that are linked to uncoupled $e$-set states. An
1072: example to this configuration is shown later in Sec.~\ref{subsec:121}.
1073: The dimension $N_D$ of the dark subspace is equal to $N_g-(N_e-N_f)$
1074: plus the dimension of the zero-subspace of the matrix $\widetilde{\bm
1075: \Pi}$, Eq.~(\ref{zero-subspace}), where we assumed that $N_g\geq
1076: (N_e-N_f)$. For $N_g< (N_e-N_f)$ the dimension of the dark subspace
1077: is equal to the dimension of the zero-subspace of the matrix
1078: $\widetilde{\bm \Pi}$.
1079:
1080: It is useful to orthogonalize the dark states of
1081: Eq.~(\ref{darkstates3}). The orthogonality relation is given by
1082: Eq.~(\ref{ortho}). In this case we find that the vectors
1083: $[\widetilde{\bm x}^{(l)}_0 \, \widetilde{\bm x}^{\prime (l)}_0]^T$,
1084: $l=1\ldots N_D$ should be the eigenvectors of the Hermitian matrix
1085: \begin{equation}\label{metric2}
1086: \left[\begin{array}{c}
1087: \widetilde{\bm P} \\ \widetilde{\bm P}'
1088: \end{array}\right]\widetilde{\bm \Sigma}^{-1\dagger}
1089: \widetilde{\bm \Sigma}^{-1} [\widetilde{{\bm P}}^\dagger
1090: \widetilde{{\bm P}}^{\prime\dagger}]\,,
1091: \end{equation}
1092: with the restriction of Eq.~(\ref{zero-subspace}).
1093:
1094:
1095: The other eigenvectors, belonging to non-zero eigenvalues, are given
1096: by
1097: \begin{equation}
1098: \widetilde{\advecb}_k \t = {1\over {\cal N}_k
1099: \t}\left[\begin{array}{c}
1100: p(t)\widetilde{{\bm P}}
1101: \widetilde{\bm y}_k \t\\ %\\
1102: p(t) \left[ \widetilde{{\bm P}}'
1103: \widetilde{\bm y}_k \t +
1104: \widetilde{{\bm \Pi}} \widetilde{\bm
1105: y}'_k \t \right]\\ %\\
1106: \eigenv_k \t \widetilde{\bm y}_k \t \\
1107: \eigenv_k \t \widetilde{\bm y}'_k \t \\ %\\
1108: s(t)\widetilde{\bm \Sigma}^{\dagger} \widetilde{\bm y}_k \t
1109: \end{array}\right],
1110: \label{eq:vk2}
1111: \end{equation}
1112: where ${\cal N}_k \t$ is a normalization factor and the vector
1113: $[\widetilde{\bm y}_k \t\, \widetilde{\bm y}'_k \t]^T$ satisfies the
1114: eigen-equation
1115: \begin{widetext}
1116: \begin{equation}
1117: \left[ \begin{array}{cc}
1118: p(t)^2(\widetilde{{\bm P}}^\dagger
1119: \widetilde{{\bm P}} +
1120: \widetilde{{\bm P}}^{\prime\dagger}
1121: \widetilde{{\bm P}}') +
1122: s(t)^2 \widetilde{\bm
1123: \Sigma}\widetilde{\bm \Sigma}^{\dagger} &
1124: p(t)^2
1125: \widetilde{{\bm P}}^{\prime\dagger} \widetilde{{\bm\Pi}} \\
1126: p(t)^2 \widetilde{{\bm\Pi}}^{\dag}\widetilde{{\bm P}}^{\prime} &
1127: p(t)^2 \widetilde{{\bm\Pi}}^{\dag}\widetilde{{\bm \Pi}}
1128: \end{array}\right] \left[ \begin{array}{c}
1129: \widetilde{\bm
1130: y}_k \t\\
1131: \widetilde{\bm y}'_k \t \end{array}\right ]
1132: = \eigenv_k \t [\eigenv_k\t-\hbar\Delta] \left[ \begin{array}{c}
1133: \widetilde{\bm y}_k \t \\ \widetilde{\bm y}'_k \t
1134: \end{array}\right ]\,.
1135: \label{eq:eigen2}
1136: \end{equation}
1137: \end{widetext}
1138: The states of Eq.~(\ref{eq:vk2}) are bright states, because they
1139: include components from the $e$ set.
1140:
1141: The eigenvectors of the Hamiltonian ${\bm H}(t)$ in the bare atomic
1142: basis can be obtained as
1143: \begin{equation}
1144: \advecb_k \t = {\bm U}^\dagger{\bm U}^{\prime\dagger} \left[
1145: \begin{array}{c}
1146: \widetilde{\bm x}_k \t \\
1147: \widetilde{\bm x}'_k \t \\
1148: \widetilde{\bm y}_k \t \\
1149: \widetilde{\bm y}'_k \t \\
1150: \widetilde{\bm z}_k \t \\
1151: \end{array} \right] =
1152: \left[ \begin{array}{c}
1153: {\bm A}^{\prime\dagger}\left[\begin{array}{c}
1154: \widetilde{\bm x}_k \t \\
1155: \widetilde{\bm x}'_k \t \\
1156: \end{array}\right] \\ \\
1157: {\bm B}^{\dagger}\left[\begin{array}{c}
1158: \widetilde{\bm y}_k \t \\
1159: {\bm B}^{\prime\dagger}\widetilde{\bm y}'_k \t \\
1160: \end{array}\right] \\ \\
1161: {\bm A}^\dagger \widetilde{\bm z}_k \t
1162: \end{array} \right]\,.
1163: \end{equation}
1164:
1165: %------------------------------------------------------
1166: \section{Adiabaticity conditions and time evolution}\label{sec:adiab-tevol}
1167: %------------------------------------------------------
1168:
1169: In Sec.~\ref{sec:mstrafo} we have presented the dark states of our
1170: degenerate system. Once we have the dark states we may consider
1171: adiabatic evolution of the system in the dark subspace. There are two
1172: questions that should be addressed in connection with adiabatic
1173: evolution:
1174: \begin{itemize}
1175: \item[(1)] What are the conditions needed to ensure adiabatic evolution?
1176: \item[(2)] If there are several degenerate dark states of a system, in
1177: general there are nonadiabatic couplings among them. How can we find
1178: the time evolution of the system in this case?
1179: \end{itemize}
1180:
1181: To answer the first question we apply the basic theory of adiabatic
1182: evolution \cite{Messiah}, which assures that the evolution is
1183: adiabatic if any nonadiabatic couplings among the adiabatic states are
1184: negligible compared with their energy separation. In our model system
1185: we have a dark subspace that is spanned by states that have eigenvalue
1186: zero. The other adiabatic states, the bright states, have non-zero
1187: eigenvalues. Because we want the state vector to remain in the dark
1188: subspace, we require that the dark subspace be separated from the
1189: bright one, as expressed by the condition
1190: \begin{equation}\label{adi-cond}
1191: \hbar |\langle \widetilde\advec^{(l)}_0(t)
1192: |\dot{\widetilde\advec}_k(t)\rangle|\ll |\eigenv_k \t|\,,
1193: \end{equation}
1194: where $l=1\ldots N_D$ and $k=1\ldots N_B$, with $N_B$ being the number
1195: of bright states. The dot denotes time derivative. We may insert into
1196: Eq.~(\ref{adi-cond}) any set of dark and bright states from
1197: Sec.~\ref{sec:mstrafo}. For example using the dark states given by
1198: Eq.~(\ref{darkstates1}) and the bright states from Eq.~(\ref{eq:vk})
1199: we find
1200: \begin{equation}\label{adi-cond2}
1201: \frac{\hbar}{{\cal N}^{(l)}_0 \t{\cal N}_k \t}
1202: \left|s(t)\dot{p}(t) - p(t)\dot{s}(t)
1203: \right| \cdot \left |\langle {\bm x}^{(l)}_0|{\bm P}
1204: |{\bm y}_k \t\rangle\right| \ll {|\eigenv_k \t|} \,.
1205: \end{equation}
1206: This formula closely resembles the adiabaticity condition of the
1207: conventional nondegenerate three-level STIRAP \cite{AAMOP},
1208: \begin{equation}
1209: \frac{1}{\Omega_0^2(t)}\left|S(t)\dot{P}(t)-P(t)\dot{S}(t)\right|\left|
1210: \begin{array}{c}
1211: \sin\varphi \\ \cos\varphi
1212: \end{array}\right|\ll \Omega_0(t)
1213: \left|\begin{array}{c}
1214: \cot\varphi \\ \tan\varphi
1215: \end{array}\right|\,,
1216: \end{equation}
1217: with $\Omega_0(t)$ and $\varphi$ defined as
1218: \begin{equation}
1219: \Omega_0(t)=\sqrt{P^2(t)+S^2(t)}\,,\qquad \tan2\varphi=\frac{\Omega_0(t)}
1220: {\Delta}\,,
1221: \end{equation}
1222: respectively. The second factor on the lhs of Eq.~(\ref{adi-cond2})
1223: contains the time-derivatives of the envelope functions of the pump
1224: and Stokes pulses, whereas the third factor involves the element of
1225: the matrix ${\bm P}$ between the $l$th dark state at the initial time
1226: and the excited-state-amplitudes of the $k$th bright state at time
1227: $t$. On the rhs $\eigenv_k$ is the eigenenergy associated with the
1228: $k$th bright state. Whenever the adiabaticity conditions
1229: Eq.~(\ref{adi-cond2}) are fulfilled for all dark and bright states of
1230: the system, then the dark and bright subspaces evolve independently.
1231:
1232:
1233: There remains the task of determining the time evolution in the dark
1234: subspace. When there are several degenerate dark states there are
1235: usually nonadiabatic couplings among them. In case of the tripod
1236: system \cite{Unanyan98,Theuer99}, due to the special choice for the
1237: time-dependence of the Stokes pulses, the two dark states mix
1238: throughout the population transfer process. If the dimension of the
1239: dark subspace is larger than two, then in general there is no exact
1240: analytic solution \cite{Kis01,Kis02}. In our case the situation is
1241: much simpler. The nonadiabatic coupling between a pair of dark states
1242: is $\langle \widetilde\advec^{(l)}_0 \t |
1243: \dot{\widetilde\advec}^{(k)}_0 \t \rangle\,$. By evaluating this
1244: expression for any pair of dark states from
1245: Sec.~\ref{sec:stokes-mstrafo} we always get identically zero. Hence
1246: the dark states do not mix throughout the whole transfer process.
1247: This property simplifies the calculations considerably, since the time
1248: evolution operator in the dark subspace is given by
1249: \begin{equation}\label{op-tevol}
1250: {\bm U}(t,t_0)=\sum_{l=1}^{N_D} |\advec^{(l)}_0(t)\rangle \langle
1251: \advec^{(l)}_0 (t_0)|\,.
1252: \end{equation}
1253: For example, for $N_g \leq N_e \leq N_f$ the vectors $\advecb^{(l)}_0
1254: (t_0)$ are equal to ${\bm x}_0^{(l)}$ from Eqs.~(\ref{darkstates1})
1255: and (\ref{ortho}). Once we define the density matrix ${\bm
1256: \varrho}(t_0)$ of the system at time $t_0$ the density matrix at any
1257: later time is given by
1258: \begin{equation}\label{rho-tevol}
1259: {\bm\varrho}(t)={\bm U}(t,t_0){\bm\varrho}(t_0){\bm U}^{\dag}(t,t_0)\,,
1260: \end{equation}
1261: provided that the adiabaticity conditions Eq.~(\ref{adi-cond2}) are
1262: satisfied. Note that this formula is valid {\em only} if
1263: ${\bm\varrho}(t_0)$ lies entirely in the dark subspace. It follows
1264: from Eqs.~(\ref{op-tevol}) and (\ref{rho-tevol}) that {\em any} pure
1265: or mixed initial state of the system occupying the dark subspace of
1266: the $g$ set is transferred to the $f$ set in the course of the
1267: population transfer process. In the case of a pure initial state
1268: ${\Psi}(t_0)$ we have
1269: \begin{equation}
1270: {\Psi}(t)={\bm U}(t,t_0){\Psi}(t_0)\,.
1271: \end{equation}
1272: Here again, ${\Psi}(t_0)$ must have components solely in the dark
1273: subspace. (Were the state vector to have components initially in the
1274: bright subspace, then adiabatic evolution would maintain such
1275: presence. Because the excited states undergo spontaneous emission,
1276: their populations have the potential to interrupt the coherence of the
1277: dynamics and thereby to diminish the population transfer.)
1278:
1279:
1280:
1281: %------------------------------------------------------
1282: \section{Some examples}\label{sec:examples}
1283: %------------------------------------------------------
1284:
1285: In this section we demonstrate through some examples the usage of our
1286: method. To be specific, we consider atomic transitions where the
1287: origin of the degeneracy is the set of degenerate magnetic sublevels
1288: of angular momentum states in the absence of a magnetic field. Our
1289: purpose is to present some typical configurations that may occur in
1290: realistic situations.
1291:
1292:
1293: %-------------------------------------------------------
1294: \subsection{The $J=1\leftrightarrow2\leftrightarrow3$ linkage}
1295: \label{subsec:123}
1296: %-------------------------------------------------------
1297:
1298: %-------------------------------------------------------
1299: \begin{figure}
1300: \includegraphics[width=7cm]{fig5} %fig5
1301: \caption{(Color Online)
1302: The coupling configuration} for the
1303: $J=1\leftrightarrow2\leftrightarrow3$ linkage with only
1304: $\sigma^{\pm}$ polarized coupling fields. The system separates
1305: into two independent subsystems; the smaller one is shown with
1306: dashed lines, the larger one with solid lines. Frame (b) shows
1307: the result of the Stokes-field MS transformation. Frame (c)
1308: shows the redefinition of the states in the $g$, $e$, and $f$
1309: sets according to Eq.~(\ref{sets}).
1310: \label{fig:123scheme}
1311: \end{figure}
1312: %-------------------------------------------------------
1313:
1314:
1315: In this example we consider the linkage $J=1 \leftrightarrow2
1316: \leftrightarrow3$, shown in Fig.~\ref{fig:123scheme} and assume that
1317: only $\sigma^{\pm}$ fields are present. In this case there are two
1318: independent coupled systems: the one with $M_g=0$, $M_e=\pm1$, and
1319: $M_f=0,\pm2$ (shown as dashed lines); and the other one with
1320: $M_g=\pm1$, $M_e=0,\pm2$, and $M_f=\pm1, \pm2$ (shown as full lines).
1321: The first one has been studied in ref.~\cite{Kis03}, hence we do not
1322: consider it here. For the second, larger system, the pump coupling
1323: matrix $\bm P$ is given by
1324: \begin{equation}\label{P234}
1325: \bm P = \frac{\hbar}{2}\sqrtfrac{1}{3}\left[ \begin{array}{ccc}
1326: \Omega_P^{(-)}&\sqrtfrac{1}{6}\Omega_P^{(+)}&0
1327: \vspace{5pt} \\
1328: 0&\sqrtfrac{1}{6}\Omega_P^{(-)}&\Omega_P^{(+)}
1329: \end{array}\right]\,,
1330: \end{equation}
1331: whereas the Stokes coupling matrix reads
1332: \begin{equation}\label{S234}
1333: {\bm S} =
1334: \frac{\hbar}{2}\sqrtfrac{1}{5}\left[
1335: \begin{array}{cccc}
1336: \vspace{5pt}
1337: \Omega_S^{(-)}&\sqrtfrac{1}{15}\Omega_S^{(+)}&0&0
1338: \\
1339: \vspace{5pt}
1340: 0&\sqrtfrac{2}{5}\Omega_S^{(-)}&\sqrtfrac{2}{5}\Omega_S^{(+)}&0
1341: \\
1342: 0&0&\sqrtfrac{1}{15}
1343: \Omega_S^{(-)}&\Omega_S^{(+)}
1344: \end{array}\right].
1345: \end{equation}
1346: The numeric factors in front of the $\Omega$-s describe the
1347: Clebsch-Gordan coefficients. The Rabi frequencies are parameterized as
1348: \begin{subequations}\label{123rabi}
1349: \begin{eqnarray}
1350: \Omega_P^{(+)}&=&\Omega_P \,e^{i\phi_P} \cos\eta\,, \\
1351: \Omega_P^{(-)}&=&\Omega_P \,e^{i\psi_P} \sin\eta\,, \\
1352: \Omega_S^{(+)}&=&\Omega_S \,e^{i\phi_S} \cos\theta\,, \\
1353: \Omega_S^{(-)}&=&\Omega_S \,e^{i\psi_S} \sin\theta\,,
1354: \end{eqnarray}
1355: \end{subequations}
1356: where the amplitudes $\Omega_{P,S}$ are nonnegative. The angles $\eta$
1357: and $\theta$ characterize the pump and Stokes field polarizations,
1358: respectively.
1359:
1360: Here $N_g < N_e < N_f$ ($2<3<4$), hence the derivation in
1361: Sec~\ref{sec:piramid1} can be applied. As a first step, we have to
1362: perform the Stokes field MS transformation. The eigenvalues
1363: $\lambda_k$ of the matrix ${\bm S}{\bm S}^{\dag}$ are given by the
1364: roots of a cubic equation, see Eq.~(\ref{eigvals123MS}).
1365: %------------------------------------------------------------------------
1366: \begin{figure}
1367: \includegraphics[width=8cm]{fig6} %fig6
1368: \caption{ (Color Online) The eigenvalues of the matrix $\bm S\bm S^{\dag}$,
1369: Eq.~(\ref{S234}), as a function of the polarization of the
1370: Stokes field. The eigenvalues are measured in the units of
1371: $(\hbar\Omega_S)^2$. }
1372: \label{fig:lambda_k}
1373: \end{figure}
1374: %------------------------------------------------------------------------
1375: We display them in Fig.~\ref{fig:lambda_k} as a function of the
1376: polarization angle $\theta$. They are never zero, hence the complete
1377: adiabatic population transfer is possible for {\em any} polarization
1378: of the Stokes field. However, their amplitudes depend on the
1379: polarization, which affects the adiabaticity conditions
1380: Eqs.~(\ref{adi-cond}) and (\ref{adi-cond2}). The Stokes field MS
1381: transformation matrices $\bm A$ and $\bm B$, Eq.~(\ref{Udef}), can be
1382: calculated in a straightforward manner; they are shown in the
1383: Appendix~\ref{sec:123MS}. Since $N_f=N_e+1$, the Stokes field MS
1384: transformation yields a transformed coupling matrix $\widetilde{\bm
1385: S}$ in the form of the first row in Eq.~(\ref{Str}). The diagonal
1386: part $\widetilde{\bm \Sigma}$ is given by
1387: \begin{equation}\label{sigma234}
1388: \widetilde{\bm\Sigma} =
1389: \sqrtfrac{7}{20}\hbar\Omega_S\left[\begin{array}{ccc}
1390: \sqrt{\lambda_1}&0&0\\
1391: 0&\sqrt{\lambda_2}&0\\
1392: 0&0&\sqrt{\lambda_3}
1393: \end{array}\right]\,.
1394: \end{equation}
1395: There are $2+4-3=3$ dark states in this system: one is in the $f$ set,
1396: an uncoupled state. The space of $g$ is two-dimensional, $N_g=2$, and
1397: hence there are two dark-states in the form of
1398: Eq.~(\ref{darkstates1}). The vectors ${\bm x}_0^{(k)}$ associated
1399: with these two dark states are the eigenvectors of the Hermitian
1400: matrix ${\bm M}$ of Eq.~(\ref{metric}) and are given in the
1401: Appendix~\ref{sec:123MS}. The two dark states are obtained by
1402: inserting their structure into Eq.~(\ref{darkstates1}) or, for the
1403: bare atomic basis, into Eq.~(\ref{advecb}).
1404:
1405: We have performed numerical simulations to check the validity of our
1406: analytic results. In Fig.~\ref{fig:123pop} initially the system was in
1407: the state $|g,J_g=1,M_g=1\rangle$. The envelope functions of the pump
1408: and Stokes pulses, respectively, are $p(t)= \exp(-[t-3]^2/6^2)$ and
1409: $s(t)= \exp(-[t+3]^2/6^2)$; and $\Omega_P=52$, $\Omega_S=42$. The
1410: polarizations are characterized by $\eta=1.3376$ rad, and
1411: $\theta=0.4636$ rad. The phases are chosen randomly as
1412: $\phi_P=1.1814$ rad, $\psi_P=0$ rad, $\phi_S=1.8925$ rad, and
1413: $\psi_S=2.8198$ rad. The detuning $\Delta$ is set to zero. We have
1414: found again very good agreement between the analytic calculations and
1415: the numeric simulation.
1416:
1417: %-------------------------------------------------------
1418: \begin{figure}
1419:
1420: \includegraphics[width=8cm]{fig7a}\\[10pt] %fig7a
1421:
1422: \includegraphics[width=8cm]{fig7b} %fig7b
1423: \caption{ (Color Online) Upper frame: the pulse sequence used
1424: for the population transfer process in the coupled angular
1425: momentum system $J=1\leftrightarrow 2\leftrightarrow 3$.
1426: Initially only the state $|g,J_g=1,M_g=1\rangle$ was populated.
1427: Lower frame: The population evolution. }
1428: \label{fig:123pop}
1429: \end{figure}
1430: %-------------------------------------------------------
1431:
1432: We also considered a mixed initial state, when the initial state of
1433: the system is chosen as half of the population is placed on each of
1434: the $|g,J_g=1,M_g=\pm1\rangle$ states, and the coherence is zero
1435: between them. The numerically calculated dynamics is shown in
1436: Fig.~\ref{fig:123-kevert-pop}. We can see that despite of the mixed
1437: initial state, the complete population can be transferred from the $g$
1438: set to the $f$ set. The pulse sequence is the same as in the previous
1439: example.
1440:
1441: %-----------------------------------------------
1442: \begin{figure}
1443: \includegraphics[width=8cm]{fig8} %fig8
1444: \caption{ (Color Online) Same as Fig.~\ref{fig:123pop}, but for a
1445: mixed initial state, $P_i(1,-1)=P_i(1,1)=1/2$, all initial
1446: coherences are zero. }
1447: \label{fig:123-kevert-pop}
1448: \end{figure}
1449: %-----------------------------------------------
1450:
1451:
1452: %-------------------------------------------------------
1453: \subsection{The $J=1\leftrightarrow1\leftrightarrow1$ linkage}
1454: \label{subsec:111}
1455: %------------------------------------------------------
1456:
1457: %-------------------------------------------------------
1458: \begin{figure}
1459: \includegraphics[width=3.5cm]{fig9} %fig9
1460: \caption{(Color Online) Same as Fig.~\ref{fig:123scheme} for
1461: equal state-degeneracies. For equal $J$-s, the
1462: $M=0\leftrightarrow 0$ transition is dipole-forbidden, and we
1463: cannot select a basis such that there occur couplings between
1464: all pairs of states of the degenerate sets. This restriction
1465: results from the property of the Clebsch-Gordan coefficients.
1466: Frame (c) shows the redefinition of the states in the $g$, $e$,
1467: and $f$ sets according to Eq.~(\ref{sets}), leading to two
1468: independent three-state linkages. }
1469: \label{fig:111scheme}
1470: \end{figure}
1471: %-------------------------------------------------------
1472:
1473:
1474: As another example we consider the linkage
1475: $J=1\leftrightarrow1\leftrightarrow1$ shown in
1476: Fig.~\ref{fig:111scheme}. In this case $N_g=N_e=N_f$, and hence the
1477: derivation in Sec.~\ref{sec:piramid1} is applicable. This is a
1478: counter-example to the general condition of Eq.~(\ref{cond}): even
1479: though the condition Eq.~(\ref{cond}) is satisfied, in this case the
1480: complete removal of an arbitrary population distribution from the $g$
1481: set is impossible in the STIRAP way.
1482:
1483: The coupling matrices $\bm S$ and $\bm P$ in the Hamiltonian
1484: Eq.~(\ref{ham}) are given by
1485: \begin{equation}
1486: \bm X = \frac{\hbar}{2}
1487: \frac{1}{\sqrt{6}}
1488: \left[ \begin{array}{ccc}
1489: \vspace{5pt}
1490: -\Omega_X^{(\pi)} &-\Omega_X^{(+)} &0
1491: \\ \vspace{5pt} [5pt]
1492: \Omega_X^{(-)}&0&-\Omega_X^{(+)}
1493: \\ [5pt]
1494: 0&
1495: \Omega_X^{(-)}&\Omega_X^{(\pi)}
1496: \end{array}\right],
1497: \end{equation}
1498: for $\bm X=\bm S$ or $\bm P$. The factor $1/\sqrt{6}$ and the $\pm$
1499: signs describe the Clebsch-Gordan coefficients. The Rabi frequencies
1500: $\Omega_X^{(\pm,\pi)}$ correspond to the $\sigma^{+}$, $\sigma^{-}$,
1501: and $\pi$ polarizations, respectively. Note that a selection rule
1502: nullifies transitions $M = 0 \leftrightarrow M = 0$.
1503:
1504: As described in Sec.~\ref{sec:stokes-mstrafo}, we perform the
1505: Stokes-field MS transformation to diagonalize the Stokes coupling
1506: matrix ${\bm S}$. The eigenvalues of the matrix ${\bm S}{\bm
1507: S}^{\dag}$ provide the squared moduli of the diagonal elements of
1508: the matrix $\widetilde{\bm \Sigma}=\widetilde{\bm S}={\bm B}{\bm
1509: S}{\bm A}^{\dag}$, Eq.~(\ref{trafo}). They are given by
1510: \begin{equation}
1511: 0,\quad \pm\frac{1}{6}\left(\Omega_{S}^{({\rm rms})}\right)^2\,,
1512: \end{equation}
1513: with $\left(\Omega_{S}^{({\rm rms})}\right)^2=|\Omega_S^{(+)}|^2 +
1514: |\Omega_S^{(-)}|^2 + |\Omega_S^{(\pi)}|^2$. One of the eigenvalues is
1515: always zero and therefore, although the system satisfies the condition
1516: for complete population transfer, Eq.~(\ref{cond}), the null Rabi
1517: frequency prevents complete transfer.
1518:
1519: Fig.~\ref{fig:111-evol} demonstrates the population transfer in this
1520: system. Initially the system was in the state
1521: $(|g,J_g=1,M_g=-1\rangle- |g,J_g=1,M_g=0\rangle +
1522: |g,J_g=1,M_g=1\rangle)/ \sqrt{3}$. The envelope functions of the pump
1523: and Stokes pulses, respectively, are chosen as
1524: $p(t)=\exp(-[t-2]^2/4^2)$ and $s(t)=\exp(-[t+2]^2/4^2)$, and
1525: $\Omega_P=\Omega_S=30$. The intensity is equally distributed among the
1526: $\sigma^{+}$, $\sigma^{-}$, and $\pi$ components of the exciting
1527: fields. The detuning $\Delta$ is set to zero. We have found
1528: excellent agreement between the analytic calculations and the numeric
1529: simulation. The adiabaticity conditions, Eq.~(\ref{adi-cond2}), are
1530: also fulfilled throughout the relevant part of the population transfer
1531: process.
1532:
1533:
1534: %-------------------------------------------------------
1535: \begin{figure}
1536:
1537: \includegraphics[width=8cm]{fig10a}\\[10pt] %fig10a
1538:
1539: \includegraphics[width=8cm]{fig10b} %fig10b
1540: \caption{ (Color Online) Upper frame: the pulse sequence used
1541: for the population transfer process in the coupled angular
1542: momentum system $J=1\leftrightarrow 1\leftrightarrow 1$.
1543: Initially the state $(|g,J_g=1,M_g=-1\rangle-
1544: |g,J_g=1,M_g=0\rangle +|g,J_g=1,M_g=1\rangle)/\sqrt{3}$ was
1545: populated. Lower frame: The population evolution. Part of the
1546: population is left in the $g$ set, because some of the MS Rabi
1547: frequencies vanish. }
1548: \label{fig:111-evol}
1549: \end{figure}
1550: %-------------------------------------------------------
1551:
1552:
1553:
1554:
1555:
1556: %---------------------------------------------------------------------
1557: \subsection{The $J=1\leftrightarrow 2 \leftrightarrow 1$ linkage}
1558: \label{subsec:121}
1559: %---------------------------------------------------------------------
1560:
1561:
1562:
1563: In our last example we consider the linkage $J=1 \leftrightarrow2
1564: \leftrightarrow1$, shown in Fig.~\ref{fig:121scheme} and assume that
1565: only $\sigma^{\pm}$ fields are present. In this case there are two
1566: independent coupled systems: the one with $M_g=0$, $M_e=\pm1$, and
1567: $M_f=0$ (shown as dashed lines) discussed recently in
1568: ref.~\cite{Shah02}, and the other one with $M_g=\pm1$, $M_e=0,\pm2$,
1569: and $M_f=\pm1$ (shown as full lines). This is a twin diamond
1570: configuration. For the larger system, the pump coupling matrix $\bm
1571: P$ is given by
1572: \begin{equation}
1573: {\bm P} = \frac{\hbar}{2}\left[ \begin{array}{ccc}
1574: \sqrtfrac{1}{3}\Omega_P^{(-)}&\sqrtfrac{1}{18}\Omega_P^{(+)}&0\\
1575: \vspace{5pt}
1576: 0&\sqrtfrac{1}{18}\Omega_P^{(-)}&\sqrtfrac{1}{3}\Omega_P^{(+)}
1577: \end{array}\right]\,.
1578: \end{equation}
1579: whereas the Stokes coupling matrix reads
1580: \begin{equation}
1581: {\bm S} = \frac{\hbar}{2}\left[ \begin{array}{cc}
1582: \frac{\sqrt{3}}{5}\Omega_S^{(-)}&0\\
1583: \vspace{5pt}
1584: \sqrtfrac{1}{50}\Omega_S^{(+)}&\sqrtfrac{1}{50}\Omega_S^{(+)}\\
1585: \vspace{5pt}
1586: 0&\frac{\sqrt{3}}{5}\Omega_S^{(-)}
1587: \end{array}\right]\,.
1588: \end{equation}
1589: The parameterization of the Rabi frequencies $\Omega_X^{(\pm)}$ is
1590: given by Eq.~(\ref{123rabi}). Here $N_g, N_f < N_e$ ($2,2<3$), hence
1591: the derivation in Sec~\ref{sec:diamond} is applicable. The sequence of
1592: the dimension of the subspaces violate the condition $N_g\leq N_e\leq
1593: N_f$, therefore, in general a STIRAP-like complete population transfer
1594: is not possible. However, this is another counter-example to the
1595: general condition of Eq.~(\ref{cond}): even though the condition
1596: Eq.~(\ref{cond}) is violated, we show that the complete removal of an
1597: arbitrary population distribution from the $g$ set is possible in the
1598: STIRAP way for a special choice of pulse polarizations and phases.
1599:
1600: As usual, we start with the Stokes field MS transformation. The
1601: eigenvalues of the matrix ${\bm S}{\bm S}^{\dag}$ are given by the
1602: roots of a quadratic equation, which read
1603: \begin{equation}\label{121seigs}
1604: \lambda_{1,2}=\frac{7}{100}\pm\frac{1}{100}\sqrt{24\cos^22\theta+1}\,.
1605: \end{equation}
1606: The Stokes field MS transformation matrices $\bm A$ and $\bm B$,
1607: Eq.~(\ref{Udef}), can be calculated in a straight forward manner, they
1608: are shown in the Appendix~\ref{sec:121MS}. Since $N_f=N_e-1$, the
1609: Stokes field MS transformation yields a transformed coupling matrix
1610: $\widetilde{\bm S}$ in the form of the last row in Eq.~(\ref{Str}).
1611: The diagonal part $\widetilde{\bm \Sigma}$ is given by
1612: \begin{equation}
1613: \widetilde{\bm\Sigma} = \frac{\hbar}{2} \Omega_S\left[\begin{array}{cc}
1614: \sqrt{\lambda_1}&0\\
1615: 0&\sqrt{\lambda_2}
1616: \end{array}\right]\,.
1617: \end{equation}
1618: The eigenvalues of Eq.~(\ref{121seigs}) are always positive, hence
1619: this matrix is nonsingular for {\em any} polarization of the Stokes
1620: field. The Stokes field MS transformation yields two $e-f$ linkages
1621: and an $e$ state which is not coupled to any $f$ state, see
1622: Fig.~\ref{fig:121scheme}b. Now, following the derivation of
1623: Sec~\ref{sec:diamond} we perform a second MS transformation for the
1624: pump field. The transformation matrix is given by Eq.~(\ref{Updef}).
1625: In our case, the $2\times 2$ unitary matrix ${\bm A}'$ is defined as
1626: \begin{widetext}
1627: \begin{equation}
1628: {\bm A}'
1629: =\left[\begin{array}{cc}
1630: \cos\theta&e^{-i(\psi_S-\phi_S)}\sin\theta\\
1631: e^{-\frac12 i(\phi_S-\psi_S+\phi_P+\psi_P)}\sin\theta&
1632: -e^{-\frac12 i(\psi_S-\phi_S+\phi_P+\psi_P)}\cos\theta
1633: \end{array}
1634: \right]\,,
1635: \end{equation}
1636: while the matrix ${\bm B}'$ is a scalar now, and chosen as unity.
1637: Since in this case $N_e-N_f<N_g$ ($3-2<2$), the transformed pump field
1638: coupling matrix takes the form of Eq.~(\ref{pi1}). The matrix
1639: $\widetilde{\bm \Pi}$ is a scalar, that reads
1640: \begin{equation}
1641: \widetilde{\bm \Pi}=\begin{array}{c}
1642: -\frac{\Omega_P}{2\sqrt{3}\sqrt{2-\cos^2 2\theta}}\left(
1643: \cos\eta\cos\theta e^{\frac12i(\psi_S-\phi_S+\phi_P-\psi_P)}
1644: -\sin\eta\sin\theta e^{-\frac12i(\psi_S-\phi_S+\phi_P-\psi_P)}\right)
1645: \end{array}\,.
1646: \end{equation}
1647: \end{widetext}
1648: This is nonzero in general, hence one of the $g$ states is linked to
1649: the {\em uncoupled} $e$ state. Therefore, there is one dark state in
1650: the system, which reads
1651: \begin{equation}\label{dark1}
1652: \widetilde{\bm \Phi}_0^{(1)}(t)=\frac{1}{{\cal N}_0^{(1)}(t)}
1653: \left[\begin{array}{c}
1654: s(t)\\
1655: 0\\
1656: 0\\
1657: 0\\
1658: 0\\
1659: -p(t)\widetilde{\bm \Sigma}^{-1}\bm B_a \bm P^\dagger {\bm A'}^\dagger
1660: \left[
1661: \begin{array}{c}1\\0\end{array}
1662: \right]
1663: \end{array}
1664: \right]\,.
1665: \end{equation}
1666: This dark state is associated with the three-state linkage in the
1667: middle of Fig.~\ref{fig:121scheme}c, indicated by heavy lines.
1668:
1669:
1670: %-------------------------------------------------------
1671: \begin{figure}
1672: \includegraphics[width=5cm]{fig11} %fig11
1673: \caption{(Color Online) Same as Fig.~\ref{fig:123scheme} for the
1674: $J=1\leftrightarrow2\leftrightarrow1$ linkage with only
1675: $\sigma^{\pm}$ polarized coupling fields. The system separates
1676: into two independent subsystems; The smaller one is shown with
1677: dashed lines, the larger one with solid lines. Frame (b) shows
1678: the result of the Stokes-field MS transformation. Frame (c)
1679: shows the result of the pump field MS transformation {\em and}
1680: redefinition of the states in the $g$, $e$, and $f$ sets
1681: according to Eq.~(\ref{sets}). In general, there is one dark
1682: state in the larger system which is associated with the
1683: three-state linkage indicated by heavy lines. }
1684: \label{fig:121scheme}
1685: \end{figure}
1686: %-------------------------------------------------------
1687:
1688:
1689:
1690: However, for
1691: \begin{subequations}\label{cptdi2}
1692: \begin{eqnarray}
1693: \psi_S-\phi_S+\phi_P-\psi_P&=&k\pi\,,\\
1694: \theta+(-1)^k\eta=\frac12\pi\,,
1695: \end{eqnarray}
1696: \end{subequations}
1697: where $k$ is an integer, the scalar $\widetilde{\bm \Pi}$ vanishes. As
1698: a result, the {\em uncoupled} $e$ state becomes decoupled from the $g$
1699: state as well. Therefore, beside the dark state of Eq.~(\ref{dark1})
1700: there is an other one
1701: \begin{equation}
1702: \widetilde{\bm \Phi}_0^{(2)}(t)=\frac{1}{{\cal N}_0^{(2)}(t)}
1703: \left[\begin{array}{c}
1704: 0\\
1705: s(t)\\
1706: 0\\
1707: 0\\
1708: 0\\
1709: -p(t)\widetilde{\bm \Sigma}^{-1}\bm B_a \bm P^\dagger {\bm A'}^\dagger
1710: \left[
1711: \begin{array}{c} 0\\1 \end{array}
1712: \right]
1713: \end{array}
1714: \right]\,.
1715: \end{equation}
1716: In summary: {\em complete} population transfer is possible from the
1717: $g$ set to the $f$ set for the special choice of pulse polarizations
1718: and phases Eq.~(\ref{cptdi2}). It is important to note that the
1719: condition for complete transfer Eq.~(\ref{cptdi2}) is equivalent to
1720: that for the diamond configuration \cite{Shah02}. Hence, the
1721: population from the total $g$ set can be transferred into the $f$ set
1722: if the condition Eq.~(\ref{cptdi2}) is fulfilled.
1723:
1724: Fig.~\ref{fig:121A-evol} demonstrates the population transfer in the
1725: twin diamond configuration. Initially the system was in the state
1726: $\cos(\alpha) |g,J_g=1,M_g=-1\rangle+ \sin(\alpha)
1727: |g,J_g=1,M_g=1\rangle$ with $\alpha=\arctan(1/3)$. The envelope
1728: functions of the pump and Stokes pulses are chosen as in
1729: Sec.~\ref{subsec:111}. The polarization of the pump and Stokes pulses
1730: were chosen as $\eta=2\pi/5$ and $\theta=-\pi/7$, respectively. All
1731: phases of the pulses are zero. The detuning $\Delta$ is set to zero.
1732: After the pulse sequence has passed, some population is left in the
1733: $g$ and $e$ sets because the polarizations of the pulses violate the
1734: special condition for complete transfer Eq.~(\ref{cptdi2}). Finally,
1735: in Fig.~\ref{fig:121B-evol} the polarizations of the pump and Stokes
1736: pulses are chosen so that the special condition Eq.~(\ref{cptdi2}) is
1737: fulfilled. Then, a complete population transfer occurs, all
1738: population from the $g$ set is moved into the $f$ set.
1739:
1740:
1741: %-------------------------------------------------------
1742: \begin{figure}
1743:
1744: \includegraphics[width=8cm]{fig12a}\\[10pt] %fig12a
1745:
1746: \includegraphics[width=8cm]{fig12b} %fig12b
1747: \caption{ (Color Online) Upper frame: the pulse sequence used
1748: for the population transfer process in the coupled angular
1749: momentum system $J=1\leftrightarrow 2\leftrightarrow 1$.
1750: Initially the state $\cos(\alpha) |g,J_g=1,M_g=-1\rangle+
1751: \sin(\alpha) |g,J_g=1,M_g=1\rangle$ with $\alpha=\arctan(1/3)$
1752: was populated. Lower frame: The population evolution. The
1753: pulses are chosen so that the special condition for complete
1754: transfer Eq.~(\ref{cptdi2}) is violated, hence part of the
1755: population is left in the $g$ and $e$ sets. }
1756: \label{fig:121A-evol}
1757: \end{figure}
1758: %-------------------------------------------------------
1759:
1760: %-------------------------------------------------------
1761: \begin{figure}
1762:
1763: \includegraphics[width=8cm]{fig13a}\\[10pt] %fig13a
1764:
1765: \includegraphics[width=8cm]{fig13b} %fig13b
1766: \caption{ (Color Online) Same as Fig.~\ref{fig:121A-evol}, but now
1767: the special condition for the pulse polarizations
1768: Eq.~(\ref{cptdi2}) is fulfilled. Upper frame: the pulse
1769: sequence used for the population transfer process. The
1770: polarizations for the pump and Stokes pulses are $\eta=2\pi/5$
1771: and $\theta=\pi/10$, respectively. Lower frame: The population
1772: evolution. All population is transferred from the $g$ set into
1773: the $f$ set. }
1774: \label{fig:121B-evol}
1775: \end{figure}
1776: %-------------------------------------------------------
1777:
1778:
1779: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1780: \section{Summary}\label{sec:summary}
1781: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1782:
1783: We have considered the extension of the well-known STIRAP process in
1784: degenerate systems in which $N_g$ degenerate states of the $g$ set are
1785: coupled by means of a pump pulse to $N_e$ degenerate states of the $e$
1786: set, which in turn are linked by the Stokes pulse to $N_f$ degenerate
1787: states of the $f$ set. We have shown that such a generalized STIRAP
1788: process is always possible if the succession of state-degeneracies is
1789: nondecreasing, i.e. $ N_g \leq N_e \leq N_f$; and the number of
1790: non-vanishing MS Rabi frequencies is at least $N_g$ for both the pump
1791: and Stokes couplings. When such conditions hold, then for arbitrary
1792: couplings among states (e.g. arbitrary elliptical polarization of
1793: electric dipole radiation between magnetic sublevels) it is possible
1794: to obtain complete adiabatic passage of all population from the states
1795: of the $g$ set into some combination of states of the $f$ set. In this
1796: process the initial state is arbitrary, it can be any pure or mixed
1797: state that occupy the $g$ set.
1798:
1799: An important exception from the above rule occurs in coupled angular
1800: momentum systems, when $J_g=J_e=J_f$. Then, due to the symmetry of
1801: the Clebsch-Gordan coefficients some couplings vanish, which results
1802: in incomplete transfer.
1803:
1804: We have examined the possibility of adiabatic passage when this
1805: restriction on degeneracies does not hold. We have shown that part of
1806: the population can be transferred to the $f$ set. We have also
1807: pointed out that, for certain choices of the polarizations of the
1808: coupling fields, complete adiabatic population transfer can be
1809: obtained.
1810:
1811: We have demonstrated that our scheme can be a powerful tool for
1812: coherent control of the quantum state in a degenerate system: in our
1813: proposal the selective addressing of individual states in the
1814: degenerate sets is not required. Nevertheless, the final state can be
1815: tailored by varying the polarizations and the relative phases of the
1816: coupling fields. We have shown through some specific examples that
1817: the control of the final superposition state is possible; the level of
1818: control depends on the system under consideration.
1819:
1820:
1821:
1822:
1823:
1824: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1825: \section*{Acknowledgments}
1826: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1827:
1828: This work has been supported by the European Union Research Training
1829: network COCOMO, contract number HPRN-CT-1999-00129. ZK and AK
1830: acknowledge the support from the Research Fund of the Hungarian
1831: Academy of Sciences (OTKA) under contract T43287. ZK acknowledges the
1832: support from the J\'anos Bolyai program of the Hungarian Academy of
1833: Sciences. He is also grateful to Prof. K. Bergmann for his kind
1834: hospitality in his group at the University of Kaiserslautern. NVV and
1835: BWS acknowledge support from the Alexander von Humboldt Foundation.
1836: BWS acknowledges support from the Graduierten Kolleg of the University
1837: of Kaiserslautern. The authors are grateful to Prof. K. Bergmann for
1838: useful discussion.
1839:
1840:
1841:
1842: \appendix
1843:
1844: %------------------------------------------------------
1845: \section{Dipole Transition Moments}
1846: %------------------------------------------------------
1847:
1848: A common situation where degeneracy occurs is when the atomic states
1849: are eigenstates of angular momentum, bearing the labels $J$ and $M$.
1850: Then the dipole moments can be expressed in terms of Clebsch-Gordan
1851: coefficients and reduced matrix elements. For the pump transition
1852: ($g-e$) the general pattern of the dipole-transition matrix elements,
1853: for arbitrary polarization, is
1854: \begin{equation}
1855: \mu_{ij} = (g|\mu|e) \sum_q \epsilon_q^{(p)}
1856: \frac{(J_g M_i, 1 q | J_e M_j)}{ \sqrt{2 J_g + 1}}\,,
1857: \quad
1858: \left\{\begin{array}{l}
1859: i=1\hdots N_g \\
1860: j=1\hdots N_e
1861: \end{array}\right.\,,
1862: \end{equation}
1863: where $(g|\mu|e)$ is the reduced matrix element and $\epsilon_q^{(p)}$
1864: parameterizes the contribution of spherical component $q$ to the
1865: interaction. The Stokes transition moments are similarly written as
1866: \begin{equation}
1867: \mu_{ij} = (e|\mu|f) \sum_q \epsilon_q^{(S)}\frac{(J_e M_i, 1 q | J_f M_j) }
1868: {\sqrt{2 J_e + 1}}\,, \quad
1869: \left\{\begin{array}{l}
1870: i=1\hdots N_e \\
1871: j=1\hdots N_f
1872: \end{array}\right..
1873: \end{equation}
1874:
1875: For vibrational transitions in molecules the reduced matrix element
1876: must include a Franck-Condon factor.
1877:
1878: %------------------------------------------------------
1879: \section{Singular coupling matrix $\widetilde{\bm\Sigma}$}
1880: \label{sec:sing-sigma}
1881: %------------------------------------------------------
1882:
1883: Let us consider the Hamiltonian Eq.~(\ref{ham}) in the MS basis,
1884: Eq.~(\ref{trafo}). The MS transformation of the coupling matrix ${\bm
1885: S}$ may result in three different forms, shown in Eq.~(\ref{Str}).
1886: We obtain a diagonal matrix $\widetilde{\bm\Sigma}$ to which are
1887: appended either rows (if $N_f < N_e$) or columns (if $N_f > N_e$) of
1888: zero values. In the discussions of Sec.~\ref{sec:mstrafo} we have
1889: assumed that the matrix $\widetilde{\bm\Sigma}$ is nonsingular. Here
1890: we consider the case when some diagonal elements of
1891: $\widetilde{\bm\Sigma}$ are zero. Let us choose the MS transformation
1892: matrices ${\bm A}$ and ${\bm B}$ in Eq.~(\ref{Udef}) in such a way
1893: that the zero diagonal elements appear in the bottom right corner of
1894: $\widetilde{\bm\Sigma}$. This non-zero part is denoted by
1895: $\widetilde{\bm\Sigma}_C$. Let the dimension of this matrix be
1896: $N_C\times N_C$. In this notation, instead of Eq.~(\ref{Str}) we have
1897: \begin{equation}\label{app:Sdef2}
1898: \widetilde{\bm S} = \left[\begin{array}{cc}
1899: \widetilde{\bm \Sigma}_C & {\bm 0} \\
1900: {\bm 0} & {\bm 0}
1901: \end{array}\right]\,,
1902: \end{equation}
1903: where the number of all zero rows is $N_e-N_C$ and the number of all
1904: zero columns is $N_f-N_C$. From this form of the Stokes coupling
1905: matrix it is clearly seen that we have $N_e-N_C$ uncoupled MS states
1906: in the $e$ set and $N_f-N_C$ uncoupled MS states in the $f$ set. By
1907: inserting the coupling matrix Eq.~(\ref{app:Sdef2}) into the
1908: transformed Hamiltonian of Eq.~(\ref{trafo}) and performing a second
1909: MS transformation as in Sec.~\ref{sec:diamond} among the $g$ set and
1910: the uncoupled MS states of the $e$ set we get
1911: \begin{equation}\label{app:Hamc}
1912: \widehat{\widetilde{{\bm H}}}(t)=\left[\begin{array}{cccccc}
1913: {\bm 0}&{\bm 0}&p(t)\widetilde{{\bm P}}&{\bm 0}&{\bm 0}&{\bm 0}\\
1914: {\bm 0}&{\bm 0}&p(t)\widetilde{{\bm
1915: P}}'&p(t)\widetilde{{\bm\Pi}}&{\bm 0}&{\bm 0}\\
1916: p(t)\widetilde{{\bm P}}^\dagger&p(t)\widetilde{{\bm P}}^{\prime\dagger}
1917: &\hbar{\bm \Delta}&{\bm 0}&s(t)\widetilde{\bm \Sigma}_C&{\bm 0}\\
1918: {\bm 0}&p(t)\widetilde{{\bm\Pi}}^{\dag}&{\bm 0}&\hbar{\bm \Delta}&{\bm 0}
1919: &{\bm 0}\\
1920: {\bm 0}&{\bm 0}&s(t)\widetilde{\bm \Sigma}_C^\dagger&{\bm
1921: 0}&{\bm 0}&{\bm 0}\\
1922: {\bm 0}&{\bm 0}&{\bm 0}&{\bm 0}&{\bm 0}&{\bm 0}
1923: \end{array}\right]\,.
1924: \end{equation}
1925: This Hamiltonian is almost identical with the one in
1926: Eq.~(\ref{Hambt}). The difference is that here on the bottom of the
1927: matrix we have some rows of zero values as well as some columns of
1928: zero values to the far right. The adiabatic states of this
1929: Hamiltonian can be found as in Sec.~\ref{sec:diamond}. The
1930: eigenvectors are parameterized as
1931: \begin{equation}
1932: \label{app:vkparam3}
1933: \widetilde{\advecb}_k = \left[\begin{array}{c}
1934: \widetilde{\bm x}_k\\
1935: \widetilde{\bm x}'_k\\
1936: \widetilde{\bm y}_k\\
1937: \widetilde{\bm y}'_k\\
1938: \widetilde{\bm z}_k\\
1939: \widetilde{\bm z}'_k\\
1940: \end{array}\right].
1941: \end{equation}
1942: The eigenvalue equation yields the set of equations as in
1943: Sec.~\ref{sec:diamond} plus one more equation for ${\bm z}'_k$
1944: \begin{equation}\label{app:extraeq}
1945: {\bm 0}=\eigenv_k{\bm z}'_k\,.
1946: \end{equation}
1947: When looking for the eigenstates belonging to the eigenvalue zero we
1948: set $\eigenv_0=0$ in Eq.~(\ref{app:extraeq}). Since ${\bm z}'_0$ does
1949: not appear in the other equations, its value is determined from the
1950: initial condition of the system. Our usual assumption is that
1951: initially only the states of the $g$ set are occupied, therefore,
1952: ${\bm z}'_0={\bm 0}$. For the eigenstates with non-zero eigenvalues
1953: the only way to satisfy Eq.~(\ref{app:extraeq}) is to set ${\bm z}'_k$
1954: to a null vector , ${\bm z}'_k={\bm 0}$. The eigenstates associated
1955: with non-zero eigenvalues $\eigenv_k$ are given in
1956: Sec.~\ref{sec:diamond}.
1957:
1958:
1959: %------------------------------------------------------
1960: \section{Linearization of the couplings $g\leftrightarrow e \leftrightarrow f$}
1961: \label{sec:linearized-couplings}
1962: %------------------------------------------------------
1963:
1964: The construction of the dark state Eq.~(\ref{darkstates1}) can be
1965: understood as follows. We introduce three sets of states, defined in
1966: the $g$, $e$, and $f$ sets, respectively
1967: \begin{subequations}\label{sets}
1968: \begin{eqnarray}
1969: \mbox{$g$ set: } \widetilde{\widetilde \psi}_g^{(l)}&=&
1970: {\bm x}_0^{(l)}\,,\quad l=1\hdots N_g\,,\label{gset}\\
1971: \mbox{$e$ set: } \widetilde{\widetilde \psi}_e^{(l)}&=&
1972: \frac{1}{{\cal N}^{(l)}_e}\widetilde{{\bm P}}^\dagger
1973: {\bm x}_0^{(l)}\,,\quad l=1\hdots N_g\,, \label{eset}\\
1974: &+& \mbox{$N_e-N_g$ other linearly independent states} \nonumber \\
1975: \mbox{$f$ set: } \widetilde{\widetilde \psi}_f^{(l)}&=&
1976: \frac{1}{{\cal N}^{(l)}_f}\widetilde{\bm \Sigma}^{-1}
1977: \widetilde{{\bm P}}^\dagger {\bm x}_0^{(l)} \,,
1978: \quad l=1\hdots N_g\,,\label{fset}\\
1979: &+& \mbox{$N_f-N_g$ other linearly independent states}\,. \nonumber
1980: \end{eqnarray}
1981: \end{subequations}
1982: The vectors ${\bm x}_0^{(l)}$ are orthonormal by construction; ${\cal
1983: N}^{(l)}_e$ and ${\cal N}^{(l)}_f$ are appropriate normalization
1984: factors for the other components. The states in the $g$ and $f$ sets
1985: of Eq.~(\ref{sets}) are orthonormal, but the states in the $e$ set of
1986: Eq.~(\ref{eset}), though linearly independent and providing a complete
1987: set of excited states, are not orthogonal. The dual counterpart
1988: \cite{dual} of the $e$ set of Eq.~(\ref{eset}) reads
1989: \begin{eqnarray}
1990: \mbox{dual $e$ set: } \widehat{\widehat \psi}_e^{(l)}&=&
1991: \frac{ {\cal N}^{(l)}_e}{{\cal N}^{(l)\,2}_f}
1992: {\bm x}_0^{(l)\,T}\widetilde{\bm P} \widetilde{\bm \Sigma}^{-1 \dag}
1993: \widetilde{\bm \Sigma}^{-1} \,,\\
1994: &&l=1\hdots N_g\,,\nonumber \\
1995: &+& \mbox{$N_e-N_g$ other linearly} \nonumber \\
1996: &&\mbox{independent states.} \nonumber
1997: \label{deset}
1998: \end{eqnarray}
1999: The vectors of these two sets are mutually orthogonal
2000: \begin{equation}
2001: \langle \widehat{\widehat \psi}_e^{(l)}| \widetilde{\widetilde \psi}_e^{(k)}
2002: \rangle=\delta_{kl}\,.
2003: \end{equation}
2004: In the basis defined by Eqs. (\ref{sets}) the Hamiltonian of
2005: Eq.~(\ref{ham-ms}) reads
2006: \begin{equation}\label{ham-northo}
2007: \widetilde{\widetilde{{\bm H}}}(t)=\left[\begin{array}{cccc}
2008: {\bm 0}&p(t)\widetilde{{\bm Q}}_1&{\bm 0}&{\bm 0} \\
2009: p(t)\widetilde{{\bm Q}}_2&\hbar{\bm \Delta} &
2010: s(t) \widetilde{\bm \Sigma}_1 & {\bm 0} \\
2011: {\bm 0}&s(t)\widetilde{\bm \Sigma}_2 & {\bm 0}&{\bm 0}\\
2012: {\bm 0}&{\bm 0}&{\bm 0}&{\bm 0}
2013: \end{array}\right]\,,
2014: \end{equation}
2015: where $\widetilde{{\bm Q}}_2$ and $\widetilde{{\bm \Sigma}}_1$ are
2016: diagonal matrices with elements
2017: \begin{subequations}
2018: \begin{eqnarray}
2019: (\widetilde{{\bm Q}}_{2})_{ll}&=&\frac{{\cal N}_f^{(l) 2}}
2020: {{\cal N}_e^{(l) }}\,,\quad l=1\hdots N_g\,,\\
2021: &\mbox{and}&\nonumber \\
2022: (\widetilde{{\bm \Sigma}}_{1})_{ll}&=&\frac{1} {{\cal N}_f^{(l) }}
2023: (\widetilde{{\bm Q}}_{2})_{ll}\,, \quad l=1\hdots N_g\,,
2024: \end{eqnarray}
2025: \end{subequations}
2026: respectively. It can be verified that the matrix elements of
2027: $\widetilde{\bm P}^{\dag}$ is zero between the rest of the dual $e$
2028: states and the $g$ states
2029: \begin{equation}
2030: \langle \widehat{\widehat \psi}_e^{(k)}|\widetilde{\bm P}^{\dag}|
2031: \widetilde{\widetilde \psi}_g^{(l)} \rangle=0\,,\quad
2032: k=N_g+1\hdots N_e\,,\,\, l=1\hdots N_g\,.
2033: \end{equation}
2034: Similarly, the matrix elements of $\widetilde{\bm \Sigma}$ is zero
2035: between the rest of the dual $e$ states and the first $N_g$ $f$ states
2036: \begin{equation}
2037: \langle \widehat{\widehat \psi}_e^{(k)}|\widetilde{\bm \Sigma}|
2038: \widetilde{\widetilde \psi}_f^{(l)} \rangle=0\,,\quad
2039: k=N_g+1\hdots N_e\,,\,\, l=1\hdots N_g\,.
2040: \end{equation}
2041: The matrix elements of the other two symmetric, non-diagonal matrices
2042: $\widetilde{{\bm Q}}_1$ and $\widetilde{{\bm \Sigma}}_2$ are given by
2043: \begin{subequations}
2044: \begin{eqnarray}
2045: (\widetilde{{\bm Q}}_{1})_{lk}&=&\frac{1} {{\cal N}_e^{(k) }}
2046: \langle {\bm x}_0^{(l)\,T}|\widetilde{\bm P} \widetilde{\bm P}^{\dag}
2047: |{\bm x}_0^{(k)}\rangle\,,
2048: \\
2049: &\mbox{and}&\nonumber \\
2050: (\widetilde{{\bm \Sigma}}_{2})_{lk}&=&\frac{1} {{\cal N}_f^{(l) }}
2051: (\widetilde{{\bm Q}}_{1})_{lk}\,.
2052: \end{eqnarray}
2053: \end{subequations}
2054: The dark states of the Hamiltonian (\ref{ham-northo}) can be obtained
2055: in the same manner as in the above derivation that led to the dark
2056: states Eq.~(\ref{darkstates1}). The population transfer is described
2057: by the equation
2058: \begin{equation}
2059: p(t){\cal N}_f^{(l) }\widetilde{x}_0^{(l)} +
2060: s(t)\widetilde{\widetilde{ z}}_0^{(l)}=0\,,
2061: \end{equation}
2062: where the components $\widetilde{x}_0^{(l)}$ and
2063: $\widetilde{\widetilde{z}}_0^{(l)}$ are the probability amplitudes
2064: associated with the basis vectors Eqs. (\ref{gset}) and (\ref{fset})
2065: in the $g$ and $f$ sets, respectively. Hence in this basis the
2066: couplings $g\leftrightarrow e \leftrightarrow f$ provide {\em
2067: independent} pathways of excitation. Each $g$ state is connected
2068: through a single pathway to a single $f$ state.
2069:
2070:
2071:
2072: \begin{widetext}
2073:
2074: %------------------------------------------------------
2075: \section{Stokes field MS transformation matrices for the
2076: $J=1\leftrightarrow2\leftrightarrow3$ linkage}\label{sec:123MS}
2077: %------------------------------------------------------
2078:
2079: The Stokes field MS transformation yields three eigenvalues
2080: $\lambda_k$ of the matrix ${\bm S}{\bm S}^{\dag}$ composed from the
2081: Stokes field coupling matrix of Eq.~(\ref{S234})
2082: \begin{equation}\label{eigvals123MS}
2083: \lambda_k = z+w\cot\left(\frac{1-k}{3}\pi+\frac{1}{3}\arctan v\right),
2084: \end{equation}
2085: for $k=1,2,3$, where
2086: \begin{subequations}
2087: \begin{eqnarray}
2088: u &=& \frac{3}{4}\sqrt{146004\cos 12\theta + 857454 \cos 8\theta
2089: + 2234532\cos 4\theta + 1524810}\\
2090: v &=& \frac{2u}{(839+909\cos 4\theta)}\\
2091: w &=& \frac{73002\cos 12\theta+428727\cos 8\theta
2092: +1117266\cos 4\theta+762405}{22960u + 19320u\cos 4\theta}\\
2093: z &=& \frac{709 \cos 4\theta + 923}{14490\cos 4\theta+17220}.
2094: \end{eqnarray}
2095: \end{subequations}
2096: The Stokes field MS transformation matrix $\bm A$ is given by
2097: \begin{equation}\label{A234}
2098: \bm A=e^{i\phi_S}\left[\begin{array}{cccc}
2099:
2100: p_1^{(A)}(\lambda_1)/n^{(A)}(\lambda_1)&p_2^{(A)}(\lambda_1)/n^{(A)}(\lambda_1)&p_3^{(A)}(\lambda_1)/n^{(A)}(\lambda_1)&p_4^{(A)}(\lambda_1)/n^{(A)}(\lambda_1)\\
2101:
2102: p_1^{(A)}(\lambda_2)/n^{(A)}(\lambda_2)&p_2^{(A)}(\lambda_2)/n^{(A)}(\lambda_2)&p_3^{(A)}(\lambda_2)/n^{(A)}(\lambda_2)&p_4^{(A)}(\lambda_2)/n^{(A)}(\lambda_2)\\
2103:
2104: p_1^{(A)}(\lambda_3)/n^{(A)}(\lambda_3)&p_2^{(A)}(\lambda_3)/n^{(A)}(\lambda_3)&p_3^{(A)}(\lambda_3)/n^{(A)}(\lambda_3)&p_4^{(A)}(\lambda_3)/n^{(A)}(\lambda_3)\\
2105:
2106: d_1^{(A)}/n_d^{(A)}&d_2^{(A)}/n_d^{(A)}&d_3^{(A)}/n_d^{(A)}&d_4^{(A)}/n_d^{(A)}
2107: \end{array}\right],
2108: \end{equation}
2109: where the polynomials $p_i^{(A)}(x)$ and the normalization $n_d^{(A)}(x)$ read
2110: \begin{subequations}
2111: \begin{eqnarray}
2112: p_1^{(A)}(x)&=&\frac18e^{i(2\psi_S-2\phi_S)}\sin\theta(14700x^2-980(2+\cos
2113: 2\theta)x+\cos4\theta+56\cos2\theta+63)\,,\\
2114: p_2^{(A)}(x)&=&\frac{\sqrt{15}}{24}e^{i(\psi_S-\phi_S)}
2115: \cos\theta(2940x^2-(308+280\cos 2\theta)x+3\cos4\theta+12\cos
2116: 2\theta+9)\,,\\
2117: p_3^{(A)}(x)&=&\frac{\sqrt{15}}{8}\sin\theta(28(1+\cos2\theta)x-
2118: \cos4\theta-4\cos2\theta-3)\,,\\
2119: p_4^{(A)}(x)&=&\frac{1}{8}e^{i(\phi_S-\psi_S)}\cos\theta(1-\cos
2120: 4\theta)\,,\\
2121: n^{(A)}(x)&=&\sqrt{ \left|p_1^{(A)}(x)\right|^2
2122: + \left|p_2^{(A)}(x)\right|^2 + \left|p_3^{(A)}(x)\right|^2 +
2123: \left|p_4^{(A)}(x)\right|^2}\,,
2124: \end{eqnarray}
2125: and the coefficients $d_i^{(A)}$, and the normalization $n_d^{(A)}$ are
2126: defined as
2127: \begin{eqnarray}
2128: d_1^{(A)}&=&-e^{i(2\psi_S-2\phi_S)}\cot^3\theta\,,\\
2129: d_2^{(A)}&=&\sqrt{15}e^{i(\psi_S-\phi_S)}\cot^2\theta\,,\\
2130: d_3^{(A)}&=&-\sqrt{15}\cot\theta\,,\\
2131: d_4^{(A)}&=&e^{i(\phi_S-\psi_S)}\,,\\
2132: n_d^{(A)}&=&\sqrt{1+15\cot^2\theta+15\cot^4\theta+\cot^6\theta}\,.
2133: \end{eqnarray}
2134: \end{subequations}
2135: Similarly, the other Stokes field MS transformation matrix is obtained
2136: as
2137: \begin{equation}\label{B234}
2138: \bm B=\left[\begin{array}{ccc}
2139: p_1^{(B)}(\lambda_1)/n^{(B)}(\lambda_1)&p_2^{(B)}(\lambda_1)/n^{(B)}(\lambda_1)&p_3^{(B)}(\lambda_1)/n^{(B)}(\lambda_1)\\
2140: p_1^{(B)}(\lambda_2)/n^{(B)}(\lambda_2)&p_2^{(B)}(\lambda_2)/n^{(B)}(\lambda_2)&p_3^{(B)}(\lambda_2)/n^{(B)}(\lambda_2)\\
2141: p_1^{(B)}(\lambda_3)/n^{(B)}(\lambda_3)&p_2^{(B)}(\lambda_3)/n^{(B)}(\lambda_3)&p_3^{(B)}(\lambda_3)/n^{(B)}(\lambda_3)
2142: \end{array}\right],
2143: \end{equation}
2144: where the polynomials $p_i^{(B)}(x)$ and the normalization
2145: $n_d^{(B)}(x)$ read
2146: \begin{subequations}
2147: \begin{eqnarray}
2148: p_1^{(B)}(x)&=&e^{i(\phi_S-\psi_S)}p_1^{(A)}(x)/\sin\theta\,,\\
2149: p_2^{(B)}(x)&=&\frac{\sqrt{6}}{12}\sin 2\theta
2150: (105 x -7 \cos 2\theta -8)\,,\\
2151: p_3^{(B)}(x)&=&\frac{1}{4}e^{i(\phi_S-\psi_S)}\sin^2 2\theta\,,\\
2152: n^{(B)}(x)&=&\sqrt{ \left|p_1^{(B)}(x)\right|^2
2153: +\left|p_2^{(B)}(x)\right|^2 +\left|p_3^{(B)}(x)\right|^2}\,.
2154: \end{eqnarray}
2155: \end{subequations}
2156: The vectors ${\bm x}_0^{(1,2)}$ characterizing the dark states of
2157: Eq.~(\ref{darkstates1}) are obtained by finding the eigenvectors of
2158: the Hermitian matrix Eq.~(\ref{metric}), which is obtained by
2159: inserting Eqs.~(\ref{P234}), (\ref{sigma234}) and (\ref{B234}) into
2160: Eq.~(\ref{metric}). The two eigenvectors are given by
2161: \begin{subequations}\label{x0:123}
2162: \begin{eqnarray}
2163: {\bm x}_0^{(1)}
2164: &=&\left[\begin{array}{c}
2165: \sin\chi e^{i\xi}\\
2166: \cos\chi
2167: \end{array}\right],\\
2168: {\bm x}_0^{(2)} &=&\left[\begin{array}{c}
2169: \cos\chi e^{i\xi}\\
2170: -\sin\chi
2171: \end{array}\right],
2172: \end{eqnarray}
2173: \end{subequations}
2174: where
2175: \begin{subequations}
2176: \begin{eqnarray}
2177: \chi&=&\frac{1}{2}\arctan\frac{2|u'|}{v'},
2178: \\
2179: \xi&=&\arg u',
2180: \\
2181: u'&=&\frac{7}{60}e^{i(\phi_S-\psi_S)}\sin2
2182: \theta(-8+7\cos2\theta\cos2\eta) \nonumber\\
2183: &&\quad +e^{i(\phi_P-\psi_P)}\sin2\eta\left[\frac{7}{24}
2184: +\left(\frac{343}{360}+\frac{7}{40}e^{2i(\phi_S-\psi_S+
2185: \psi_P-\phi_P)}\right) \sin^22\theta\right],
2186: \\
2187: v'&=&\frac{49}{60}\cos(\phi_S-\psi_S+\psi_P-\phi_P)
2188: \sin2\eta\sin4\theta +\left(\frac{301}{36}+
2189: \frac{203}{90}\cos^22\theta\right)\cos2\eta-\frac{49}{5}\cos2\theta.
2190: \end{eqnarray}
2191: \end{subequations}
2192:
2193:
2194:
2195:
2196:
2197: %------------------------------------------------------
2198: \section{Stokes field MS transformation matrices for the
2199: $J=1\leftrightarrow2\leftrightarrow1$ linkage}\label{sec:121MS}
2200: %------------------------------------------------------
2201:
2202:
2203: The Stokes field MS transformation matrix $\bm A$ is given by
2204: \begin{equation}
2205: \bm A = e^{i\psi_S}\left[\begin{array}{cc}
2206: p_1^{(A)}(\lambda_1)/n^{(A)}(\lambda_1)&p_2^{(A)}(\lambda_1)/
2207: n^{(A)}(\lambda_1)\\
2208: p_1^{(A)}(\lambda_2)/n^{(A)}(\lambda_2)&p_2^{(A)}(\lambda_2)/
2209: n^{(A)}(\lambda_2)
2210: \end{array}\right]\,,
2211: \end{equation}
2212: where the polynomials $p_i^{(A)}(x)$ and the normalization $n^{(A)}(x)$ read
2213: \begin{eqnarray}
2214: p_1^{(A)}(x)&=&-1-5\sin^2\theta+50 x,\\
2215: p_2^{(A)}(x)&=&\sin\theta\cos\theta e^{-i(\psi_S-\phi_S)},\\
2216: n^{(A)}(x)&=&\sqrt{ \left|p_1^{(A)}(x)\right|^2 +
2217: \left|p_2^{(A)}(x)\right|^2}.
2218: \end{eqnarray}
2219: Similarly, the other Stokes field MS transformation matrix is obtained
2220: as
2221: \begin{equation}
2222: \bm B = \left[\begin{array}{c}
2223: \bm B_a\\
2224: \bm B_b
2225: \end{array}\right]\,,
2226: \end{equation}
2227: where
2228: \begin{equation}
2229: \bm B_a = \left[\begin{array}{cc}
2230: \mathrm{sgn}(\sin4\theta\sin\theta)&0\\
2231: 0&\mathrm{sgn}(\cos \theta)
2232: \end{array}\right]
2233: \left[\begin{array}{ccc}
2234: p_1^{(B)}(\lambda_1)/n^{(B)}(\lambda_1)&p_2^{(B)}
2235: (\lambda_1)/n^{(B)}(\lambda_1)&p_3^{(B)}(\lambda_1)/n^{(B)}(\lambda_1)\\
2236: p_1^{(B)}(\lambda_2)/n^{(B)}(\lambda_2)&p_2^{(B)}(\lambda_2)/n^{(B)}
2237: (\lambda_2)&p_3^{(B)}(\lambda_2)/n^{(B)}(\lambda_2)
2238: \end{array}\right]\,,
2239: \end{equation}
2240: and
2241: \begin{equation}
2242: \bm B_b = \left[\begin{array}{ccc}
2243: d_1^{(B)}/n_d^{(B)}&d_2^{(B)}/n_d^{(B)}&d_3^{(B)}/n_d^{(B)}
2244: \end{array}\right]\,.
2245: \end{equation}
2246: The polynomials $p_i^{(B)}(x)$ with the normalization $n_d^{(B)}(x)$;
2247: the coefficients $d_i^{(B)}$ with the normalization $n_d^{(B)}$ read
2248: \begin{eqnarray}
2249: p_1^{(B)}(x)&=&-e^{-i(\phi_S-\psi_S)} \cos^2\theta
2250: (7\sin^2\theta+\cos^2\theta-50x)\,,\\
2251: p_2^{(B)}(x)&=&\frac{\sqrt{6}}{4}\sin4\theta\,,\\
2252: p_3^{(B)}(x)&=&e^{-i(\psi_S-\phi_S)} \sin^2\theta
2253: (7\cos^2\theta+sin^2\theta-50x)\,,\\
2254: n^{(B)}(x)&=&\sqrt{ \left|p_1^{(A)}(x)\right|^2 +
2255: \left|p_2^{(A)}(x)\right|^2 + \left|p_3^{(A)}(x)\right|^2}\,,\\
2256: d_1^{(B)}&=&e^{i(\phi_S-\psi_S)}\sin^2\theta\,,\\
2257: d_2^{(B)}&=&-\sqrt{6}\sin\theta\cos\theta\,,\\
2258: d_3^{(B)}&=&e^{i(\psi_S-\phi_S)}\cos^2\theta\,,\\
2259: n_d^{(B)}&=&\sqrt{1+\sin^2 2\theta}\,.
2260: \end{eqnarray}
2261:
2262:
2263:
2264: \end{widetext}
2265:
2266:
2267:
2268: \begin{thebibliography}{99}
2269:
2270: \bibitem{Vitanov01} N.V. Vitanov, M. Fleischhauer, B.W. Shore, and K.
2271: Bergmann, Adv. Atomic Mol. Opt. Phys. {\bf 46}, 55 (2001).
2272:
2273: \bibitem{STIRAP} J. Oreg, F. T. Hioe, J.H. Eberly, Phys. Rev. A {\bf
2274: 29}, 690 (1984); J. R. Kuklinski, U. Gaubatz, F. T. Hioe, and
2275: K. Bergmann, Phys. Rev. A {\bf 40}, 6741 (1989); U. Gaubatz, P.
2276: Rudecki, S. Schiemann, and K. Bergmann, J. Chem. Phys. {\bf 92},
2277: 5363 (1990).
2278:
2279: \bibitem{AAMOP} N.V. Vitanov, T. Halfmann, B.W. Shore, and K.
2280: Bergmann, Ann. Rev. Phys. Chem. {\bf 52}, 763 (2001)
2281:
2282: \bibitem{Shore91} B.W. Shore, K. Bergmann, J. Oreg, and S. Resenwaks,
2283: Phys. Rev. A {\bf 44}, 7442 (1991).
2284:
2285: \bibitem{Smith92} A.V. Smith, J. Opt. Soc. Am. B {\bf 9}, 1543 (1992).
2286:
2287: \bibitem{Pillet93} P. Pillet, C. Valentin, R.-L. Yuan, and J. Yu,
2288: Phys. Rev. A {\bf 48}, 845 (1993).
2289:
2290: \bibitem{Weiss94} D.S. Weiss, B.C. Young, S. Chu, Appl. Phys. B {\bf
2291: 59}, 217 (1994).
2292:
2293: \bibitem{Shore95} B.W. Shore, J. Martin, M.P. Fewell, K. Bergmann,
2294: Phys. Rev. A {\bf 52}, 566 (1995).
2295:
2296: \bibitem{Martin95} J. Martin, B. W. Shore, and K. Bergmann, Phys. Rev.
2297: A {\bf 52}, 583 (1995).
2298:
2299: \bibitem{Malinovsky97} V.S. Malinovsky, D.J. Tannor, Phys. Rev. A
2300: {\bf 56}, 4929 (1997).
2301:
2302: \bibitem{Vitanov98} N.V. Vitanov, Phys. Rev. A {\bf 58}, 2295 (1998).
2303:
2304: \bibitem{Theuer98} H. Theuer and K. Bergmann, Eur. Phys. J. D {\bf
2305: 2}, 279 (1998).
2306:
2307: % Coherent superposition
2308: \bibitem{Marte91} P.~Marte, P.~Zoller and J.L.~Hall, Phys. Rev. A {\bf
2309: 44}, R4118 (1991).
2310:
2311: \bibitem{Lawall94} J. Lawall and M. Prentiss, Phys. Rev. Lett. {\bf
2312: 72}, 993 (1994).
2313:
2314: \bibitem{Goldner94} L.S. Goldner, C. Gerz, R.J.C. Spreeuw, S.L.
2315: Rolston, C.I. Westbrook, W.D. Phillips, P. Marte, and P. Zoller,
2316: Phys. Rev. Lett. {\bf 72}, 997 (1994).
2317:
2318: \bibitem{Weitz94} M. Weitz, B.C. Young, and S. Chu, Phys. Rev. Lett.
2319: {\bf 73} 2563 (1994).
2320:
2321: % Tripod: theory
2322: \bibitem{Unanyan98} R.G. Unanyan, M. Fleischhauer, B.W. Shore, and K.
2323: Bergmann, Opt. Commun. {\bf 155}, 144 (1998).
2324:
2325: % Tripod: experiment
2326: \bibitem{Theuer99} H. Theuer, R.G. Unanyan, C. Habscheid, K. Klein,
2327: and K. Bergmann, Opt. Express {\bf 4}, 77 (1999).
2328:
2329: \bibitem{Unanyan99} R.G. Unanyan, B.W. Shore, and K. Bergmann, Phys.
2330: Rev. A {\bf 59}, 2910 (1999).
2331:
2332: % Preparation of an N-component maximal coherent superposition state ...
2333: \bibitem{Unanyan01} R.G. Unanyan, B.W. Shore, and K. Bergmann, Phys.
2334: Rev. A {\bf 63}, 043401 (2001).
2335:
2336: % Coherent superposition state in Degenerate STIRAP
2337: \bibitem{Kis01} Z. Kis and S. Stenholm, Phys. Rev. A {\bf 64}, 063406
2338: (2001).
2339:
2340: \bibitem{Kis02} Z. Kis and S. Stenholm, J. Mod. Optics {\bf 49}, 111
2341: (2002).
2342:
2343: \bibitem{Kraal02a} P. Kr\'al, Z. Amitay, M. Shapiro, Phys. Rev. Lett.
2344: {\bf 89}, 63002 (2002).
2345:
2346: \bibitem{Kraal02b} P. Kr\'al, M. Shapiro, Phys. Rev. A {\bf 65}, 43413
2347: (2002).
2348:
2349: \bibitem{Kis03} A. Karpati and Z. Kis, J. Phys. B {\bf 36}, 905
2350: (2003).
2351:
2352: \bibitem{Morris83} J. R. Morris and B. W. Shore, Phys. Rev. A {\bf
2353: 27}, 906 (1983).
2354:
2355: \bibitem{5ss} N.V. Vitanov, Z. Kis, and B. W. Shore, Phys. Rev. A {\bf
2356: 68}, 063414 (2003).
2357:
2358:
2359: \bibitem{dual} S. Lipschutz, M.L. Lipson, {\em Schaum's Outline of
2360: Linear Algebra} (McGraw-Hill).
2361:
2362: \bibitem{Messiah} A. Messiah, 1959, {\em M\'ecanique Quantique}
2363: (Paris: Dunod) pp.637-650.
2364:
2365: \bibitem{Shah02} S.P. Shah, D.J. Tannor, and S.A. Rice, Phys. Rev. A
2366: {\bf 66}, 033405 (2002).
2367:
2368:
2369:
2370:
2371:
2372: \end{thebibliography}
2373: \end{document}
2374: