0705.1299/pra.tex
1:  \documentclass[aps,showpacs,amsmath,amssymb,pra,twocolumn,superscriptaddress]{revtex4}
2: \usepackage{graphicx}% Include figure files
3: \usepackage{bm}% bold math
4: \usepackage{color}
5: \definecolor{mygrey}{rgb}{.9,.9,.9}
6: \hyphenation{Ryd-berg con-figu-ra-tion}
7: 
8: \begin{document}
9: \title{Ultracold Rydberg Atoms in a Ioffe-Pritchard Trap}
10: \date{\today}
11: 
12: \author{Bernd Hezel}
13: \email[]{hezel@physi.uni-heidelberg.de}
14: \affiliation{Physikalisches Institut, Universit\"at Heidelberg, Philosophenweg 12, 69120 Heidelberg, Germany}
15: 
16: \author{Igor Lesanovsky}
17: \email[]{Igor.Lesanovsky@uibk.ac.at}
18: \affiliation{Institut f\"ur Theoretische Physik, Universit\"at Innsbruck, Technikerstr. 25, 6020 Innsbruck, Austria}
19: 
20: \author{Peter Schmelcher}
21: \email[]{Peter.Schmelcher@pci.uni-heidelberg.de}
22: \affiliation{Physikalisches Institut, Universit\"at Heidelberg, Philosophenweg 12, 69120 Heidelberg, Germany}
23: \affiliation{Theoretische Chemie, Physikalisch-Chemisches Institut, Universit\"at Heidelberg, INF 229, 69120 Heidelberg, Germany}%
24: 
25: \begin{abstract}
26: We discuss the properties of ultracold Rydberg atoms in a Ioffe-Pritchard magnetic field configuration. 
27: The derived two-body Hamiltonian unveils how the large size of Rydberg atoms affects their coupling to the inhomogeneous magnetic field. 
28: The properties of the compound electronic and center of mass quantum states are thoroughly analyzed. 
29: We find very tight confinement of the center of mass motion in two dimensions to be achievable  
30: while barely changing the electronic structure compared to the field free case. 
31: This paves the way for generating a one-dimensional ultracold quantum Rydberg gas. 
32: \end{abstract}
33: 
34: \pacs{32.60.+i,33.55.Be,32.10.Dk,33.80.Ps}
35: \maketitle
36: 
37:  \section{\label{s:intro}Introduction}
38: 
39: Powerful experimental cooling techniques have been developed in the past decades that
40: allow us to probe the micro and nanokelvin regime while controlling 
41: the internal and external degrees of freedom of atomic systems.
42: As a result dilute ultracold gases that qualify perfectly for 
43: the study of quantum phenomena on a macroscopic scale \cite{pethick01,dalfovo99,pitaevskii03}
44: can nowadays be prepared almost routinely.
45: Although being dilute, interactions play an important role and
46: rich collective phenomena, reminiscent of e.g. those in traditional condensed matter physics, appear.
47: 
48: The attractiveness of Rydberg atoms arises from their extraordinary properties \cite{gallagher}.
49: The large displacement of the valence electron and the atomic core 
50: is responsible for the exaggerated response to external fields and, therewith, for their enormous polarizability.
51: Rydberg atoms possess large dipole moments and, despite being electronically highly excited, 
52: they can possess lifetimes of the order of milliseconds or even more.
53: 
54: Due to their susceptibility with respect to external fields and/or their long range
55: interaction, ensembles of Rydberg atoms represent intriguing many-body systems with
56: rich excitations and decay channels. Starting from laser cooled ground state atoms, a laser typically
57: excites a subensemble of the atoms to the desired Rydberg states.
58: Since the ultraslow motion of the atoms can be ignored on short timescales, 
59: Rydberg-Rydberg interactions dominate the system and we encounter a so-called frozen
60: Rydberg gas \cite{mourachko}. The strength of the interaction can be varied by tuning external fields
61: and/or by selecting specific atomic states. An exciting objective are the many-body effects
62: to be unraveled in ultracold Rydberg gases (see Refs. \cite{pohl1,pohl2} and references therein).
63: At a certain stage of the evolution ionization might take over leading to a cold Rydberg plasma. 
64: 
65: Beyond the above there is a number of topical and promising research activities involving
66: cold Rydberg states. One example are long range molecular Rydberg states \cite{greene}
67: with unusual properties if exposed to magnetic fields \cite{igor:jpb39}. 
68: Another one is due to the strong dipole-dipole interaction of Rydberg atoms which strongly inhibits excitation
69: of their neighbors \cite{singer,tong}.  The resulting local excitation blockade is state dependent 
70: and can turn Rydberg atoms into possible candidates for 
71: quantum information processing schemes~\cite{ryabtsev,lukin}. 
72:  
73: A precondition for enabling the processing of Rydberg atoms 
74: is the availability of tools to control their quantum behavior and properties.
75: An essential ingredient in this respect is the trapping of electronically highly excited atoms.
76: The present work provides a major contribution on this score. Let us briefly address
77: previous works on Rydberg atoms exposed to inhomogeneous static field configurations.
78: 
79: First evidence for trapped Rydberg gases has been experimentally found by Choi~et~al.~\cite{choi,choi:epjd}. 
80: The authors use strong bias fields to trap ``guiding center'' drift atoms for up to $200$~ms. 
81: Quantum mechanical studies of highly excited atoms in magnetic quadrupole fields demonstrated
82: the existence of e.g. intriguing spin polarization patterns and magnetic field-induced
83: electric dipole moments \cite{igor:epl,igor:jpb}. These investigations were based on the 
84: assumption of an infinitely heavy nucleus. A description of the coupled center of mass (c.m.) and
85: electronic dynamics has been presented in Refs.\cite{igor:prl,igor:pra72}: 
86: Trapping has been achieved for quantum states with sufficiently large total, i.e.\ electronic
87: and c.m., angular momentum.  Pictorially speaking this addresses atoms that circle
88: around the point of zero field at a sufficiently large distance.
89: Recently it has been demonstrated that trapping in a Ioffe-Pritchard configuration
90: is possible without imposing the condition of large c.m. angular momenta \cite{hezel:prl}. 
91: The present investigation works out this setup in detail and provides comprehensive results
92: for Rydberg atoms exposed to the Ioffe-Pritchard field configuration.
93: 
94: In detail we proceed as follows.
95: Sect.~\ref{s:h} contains a derivation of our working Hamiltonian for a highly excited 
96: atom in the inhomogeneous field including the coupling of the electronic and c.m.\ motion of the atom. 
97: In Sect.~\ref{s:aa} we introduce an adiabatic approximation in order to solve the corresponding
98: stationary Schr\"odinger equation. In Sect.~\ref{s:ees} we analyze the obtained spectra and
99: point out the capacity of the Ioffe bias field 
100: to regulate the distance between the surfaces, and with that the quality of the adiabatic approach. 
101: Intersections through the surfaces show their deformation when the field gradient is increased. 
102: Subsequently we characterize the electronic wave functions by discussing relevant expectation values.
103: Sect.~\ref{s:cm} is dedicated to the c.m. dynamics in the uppermost adiabatic energy surface.
104: We arrive at a confined quantized c.m. motion without the need to impose any restriction
105: on its properties. Examining the fully quantized states we observe that 
106: the extension of the electronic cloud can exceed the extension of the c.m.\ wave function. 
107: 
108: \section{\label{s:h}Hamiltonian}
109: 
110: \subsection{\label{s:h:tbh}Two-body Approach}
111: 
112: The large distance of the highly excited
113: valence electron (particle 1) from the remaining closed-shell ionic core of an alkali Rydberg atom (particle 2)
114: renders it possible to model the mutual interaction by an effective potential 
115: which is assumed to depend only on the distance of the two particles.
116: For alkali atoms, in particular, 
117: whose core possess zero total angular momentum and zero total spin, 
118: the only essential difference to the Coulombic case is due to the finite size of the core.
119: In any case, the effective potential $V(r)$ only noticeably differs 
120: from the pure Coulomb potential at small distances~$r$.
121: States of high \emph{electronic} angular momenta $l$, 
122: on which we focus in the present investigation, 
123: almost exclusively probe the Coulombic tail of this potential.
124: 
125: The coupling of the charged particles to the external magnetic field 
126: is introduced via the minimal coupling, 
127: $\bm{p}\rightarrow\bm{p}-q \bm{A}$, where $q$ is the charge of the particle and $\bm{A}$ is a vector potential 
128: belonging to the magnetic field $\bm{B}$.  Including the coupling of the magnetic moments to the external field 
129: ($\bm{\mu}_1$ and $\bm{\mu}_2$ originate from the electronic and nuclear spin, respectively), 
130: our initial Hamiltonian reads 
131: (we use atomic units except when stated otherwise) 
132: \begin{eqnarray} \label{eq:Hinit}
133:     H_{init}&=&\frac{1}{2M_1}\left(\bm{p}_1-q_1\bm{A}(\bm{r}_1)\right)^2
134:     +\frac{1}{2M_2}(\bm{p}_2-q_2\bm{A}(\bm{r}_2))^2 \nonumber\\
135:     &&+V(\left|\bm{r}_1-\bm{r}_2\right|)
136:     -\bm{\mu}_1\cdot\bm{B}(\bm{r}_1)-\bm{\mu}_2\cdot\bm{B}(\bm{r}_2) \; .
137: \end{eqnarray}
138: We do not take into account spin-orbit-coupling and relativistic mass changes. 
139: The difference in energy shift for adjacent, large angular momentum states ($l$, $l\pm 1$)  
140: due to these relativistic corrections 
141: is $\Delta W_{FS}=\alpha^2/2n^5$ \cite{bethe}, 
142: where $\alpha$ is the fine structure constant, and therefore negligible for Rydberg states.
143: At $n=30$ one receives $\Delta W_FS=1.1\times 10^{-12}$ atomic units.
144: To give an idea of the scope of this approximation we anticipate a result from Sec.~\ref{s:ees}: 
145: The energy gap between two adjacent high-l electronic states is approximately $E_{dist}=B/2$ a.u. 
146: Demanding $\Delta W_{FS} / E_{dist} \ll 1$ results is constraining the Ioffe field strength $B$ 
147: to be much larger than $5$ mG.
148: 
149: Before we focus on the Ioffe-Pritchard configuration 
150: let us first examine a general field  $\bm B$ composed of 
151: a constant term $\bm B_c$, a linear term $\bm B_l$ and higher order terms, $\bm{B}=\sum \bm{B}_i$.
152: The vector potential shall satisfy the Coulomb gauge. 
153: The squared terms can then be simplified taking advantage of the vanishing commutator 
154: $[\bm{A}(\bm{r}_1),\bm{p}_1]$ to obtain 
155: $(\bm{p}_1-q \bm{A}(\bm{r}_1))^2= \bm{p}_1^2-2 q \bm{A}(\bm{r}_1)\cdot\bm{p}_1+ q^2 \bm{A}(\bm{r}_1)^2$.
156: In the so-called symmetric gauge 
157: the vector potential of a constant magnetic field is given by
158: $\bm{A}_c(\bm{r}_1)=1/2\: \bm{B}_c\times\bm{r}_1$. 
159: The analogon for a linear field is 
160: $\bm{A}_l(\bm{r}_1)=1/3\: \bm{B}_l(\bm{r}_1)\times\bm{r}_1$.
161: It can be proven that the vector potential of an arbitrary magnetic field can be expanded
162: in a corresponding form \cite{igor:diss}
163: permitting a representation of the vector potential as a cross product 
164: $\bm{A}(\bm{r}_1) = \sum_i \bm{A}_i(\bm{r}_1) 
165: =\bm{\tilde{B}}(\bm{r}_1)\times\bm{r}_1$,
166: where $\bm{\tilde{B}}(\bm{r}_1)=\sum g_i \bm{B}_i(\bm{r}_1)$ 
167: and $i\in \{c,l,\dots\}$ denotes the order of the corresponding terms of $\bm A$ and $\bm B$ 
168: with respect to spacial coordinates. 
169: $g_i$ are the coefficients $\frac{1}{2}$, $\frac{1}{3}$ etc.
170: The particular form of this potential and the vanishing divergence of magnetic fields 
171: admit the simplification
172: \begin{equation}                                                                 \label{eq:simpleAp}
173: \bm{A}(\bm{r}_1)\cdot\bm{p}_1=(\bm{r}_1\times\bm{p}_1)\cdot\bm{\tilde{B}}(\bm{r}_1)
174: =\bm{L}_1\cdot\bm{\tilde{B}}(\bm{r}_1) \quad , 
175: \end{equation}
176: where we exemplarily defined the angular momentum of particle 1, $\bm{L}_1=\bm{r}_1\times\bm{p}_1$.
177: 
178: Since the interaction potential depends only on the distance of the two particles, 
179: it is natural to introduce relative and c.m. coordinates, 
180: $\bm r_1=\bm R+(M_2/M)\bm r$ and 
181: $\bm r_2=\bm R-(M_1/M)\bm r$
182: with the total mass $M=M_1+M_2$.
183: If no external field was present, the new coordinates would decouple 
184: the internal degrees of freedom from the external c.m.\ ones. 
185: Yet even a homogeneous magnetic field 
186: couples the relative and the c.m.~motion \cite{dippel,peter:pla}. 
187: For neutral systems in static homogeneous magnetic fields, however, 
188: a so-called `pseudoseparation' can be performed providing us with an effective Hamiltonian for the relative motion, 
189: that depends on the c.m. motion only parametrically via the eigenvalues of the pseudomomentum
190: \cite{avron,peter:cpl208,peter:ctqc,dippel} which is associated with the c.m.\ motion.
191: Such a procedure is not available in the present case of a more general inhomogeneous field.
192: In the new coordinate system the Hamiltonian (\ref{eq:Hinit}) becomes
193: \begin{multline}
194:   H= H_0 
195:   + \bm{L}_1 \bm{\tilde{B}}(\bm{R}+\frac{M_2}{M}\bm{r})
196:   - \bm{L}_2 \bm{\tilde{B}}(\bm{R}-\frac{M_1}{M}\bm{r}) \\
197:   - \bm{\mu}_1\bm{B}(\bm R+\frac{M_2}{M}\bm r)
198:   - \bm{\mu}_2\bm{B}(\bm R-\frac{M_1}{M}\bm r) 
199:   + \mathcal O(\bm A^2)\ ,
200: \end{multline}
201: where the angular momenta of the particles read
202: \begin{align}
203:   \bm{L}_{1} = ({M_{1}}/{M})\bm{L}_R 
204:   + ({M_{2}}/{M})\bm{L}_r 
205:   + \bm{R}\times \bm{p} 
206:   + ({m}/{M}) \bm{r}\times\bm{P} \nonumber \\
207:   \bm{L}_{2} = ({M_{2}}/{M})\bm{L}_R 
208:   + ({M_{1}}/{M})\bm{L}_r 
209:   - \bm{R}\times \bm{p} 
210:   - ({m}/{M}) \bm{r}\times\bm{P} \nonumber 
211: \end{align}
212: (see also Ref.\ \cite{igor:pra72}), and the terms that do not depend on the field are summarized to 
213: $H_0=\frac{\bm{p}^2}{2m}+\frac{\bm{P}^2}{2M}+V(\bm{r})$. 
214: Here, $\bm{L}_{\bm{r}}=\bm{r}\times\bm{p}$, $\bm{L}_{\bm{R}}=\bm{R}\times\bm{P}$, and the reduced mass $m=M_1 M_2/M$ have been introduced. 
215: 
216: To simplify the Hamiltonian we apply a unitary transformation 
217: that eliminates c.m.~momentum dependent coupling terms generated by the homogeneous field component
218: \begin{equation}                                                                  \label{eq:U}
219:   U=\exp\left\{\frac{i}{2}\,\bm{B}_c\times \bm{r}\cdot\bm{R}\right\} \; .
220: \end{equation}
221: $H_0$ transforms as follows 
222: \[  U^{\dagger}H_0 U= 
223:   H_0+\frac{1}{2}\bm{B}_c \left(-\frac{1}{m}\bm{R}\times \bm{p}+\frac{1}{M}\bm{r}\times\bm{P}\right)
224:   +\mathcal{O}(\bm{B}_c^2).\]
225: The transformation of the remaining terms generates exclusively additional terms, that are quadratic
226: with respect to the magnetic field. 
227: Exploiting now the fact that the mass of the ionic core is much larger than the mass of the valence electron, 
228: we only keep magnetic field dependent terms of the order of the inverse light mass $1/M_1$ 
229: (which becomes $1$ in atomic units).
230: We arrive at the Hamiltonian 
231: \begin{multline}                                                                  \label{eq:UHU}
232:   U^\dagger H U     = {\bm{p}^2}/{2} + U^{\dagger}V(\bm{r})U + {\bm{P}^2}/{2M} 
233:    + {1}/{2}\;\bm{L}_{\bm r}\cdot\bm{B}_c \\
234:    + \bm{A}_l(\bm{R}+\bm{r})\cdot\bm{p}                             
235:    +(\bm{L}_{\bm r} +\bm{R}\times\bm{p})\cdot\bm{\tilde{B}}_{n}(\bm R + \bm r) \\
236:    - \bm{\mu}_1\cdot\bm B(\bm R + \bm r)-\bm{\mu}_2\cdot \bm{B}(\bm R) \; .    
237: \end{multline}
238: 
239: The diamagnetic terms which are proportional to $\bm{A}^2$ 
240: (and herewith proportional to $\bm{B}^2$, see Eq.~(\ref{eq:simpleAp}))
241: have been neglected. 
242: Due to the unitary transformation $U$, $\bm R$-dependent terms 
243: that are quadratic in the Ioffe field strength $B$ do not occur 
244: and only an electronic term $B^2 (x^2+y^2)/8$ remains 
245: whose typical energy contribution amounts to $B^2 n^4/8\approx 10^{5} B^2$ for $n=30$. 
246: Besides we obtain a term quadratic in the field gradient $G$. 
247: The term quadratic in the Ioffe field is negligible 
248: in comparison with the dominant shift due to the linear Zeeman term 
249: as long as $B$ is significantly smaller than $10^4$~Gauss which is guaranteed in our case. 
250: Moreover, the c.~m.\ coordinate dependance of this diamagnetic term 
251: is much weaker than the c.~m.\ coordinate dependance of the terms linear in the field gradient. 
252: The term quadratic in the field gradient can be neglected in comparison with the corresponding linear term.
253: Up to now we did not use the explicit form of the Ioffe-Pritchard field configuration.
254: (In anticipation of the special field configuration we leave the term containing $\bm A_l$ in its original form.)
255: 
256: 
257: 
258: \subsection{\label{ch:ip}Ioffe-Pritchard Field Configuration}
259: 
260: %%############################################
261: Two widely spread magnetic field configurations that exhibit a local field minimum and serve
262: as key ingredients for the trapping of weak-field seeking atoms
263: are the 3D quadrupole and the Ioffe-Pritchard configuration.
264: The Ioffe-Pritchard configuration resolves the problem of particle loss due to spin flip
265: by means of an additional constant magnetic field.
266: A macroscopic realization uses four parallel current carrying Ioffe bars which generate the quadrupole field.
267: Encompassing Helmholtz coils create the additional constant field. There are many alternative layouts, 
268: the field of a clover-leaf trap for example features the same expansion around the origin \cite{cloverleaf}. 
269: On a microscopic scale the Ioffe-Pritchard trap has been implemented on atom chips by a $Z$-shaped wire \cite{folman}.  
270: 
271: The vector potential and the magnetic field read
272: \begin{gather}  
273:   \label{eq:overallpotential}
274:   \bm{A}=\underbrace{ \frac{B}{2} \left( \begin{array}{ccc} -y \\ x \\ 0 \end{array} \right)}_{=\bm{A}_c}
275:   +\underbrace{G\left( \begin{array}{ccc} 0 \\ 0 \\ x y \end{array} \right)}_{=\bm{A}_l}
276:   +\bm{A}_q \; , \\
277:   \label{eq:overallfield}
278:   \bm{B}= \underbrace{B \left( \begin{array}{ccc} 0 \\ 0 \\ 1 \end{array} \right)}_{=\bm{B}_c}
279:   +\underbrace{G \left( \begin{array}{ccc} x \\ -y \\ 0 \end{array} \right)}_{=\bm{B}_l}
280:   +\bm{B}_q \; .
281: \end{gather}
282: where 
283: $\bm{A}_q=\frac{Q}{4}(x^2+y^2-4 z^2) (-y \mathbf{e}_x + x \mathbf{e}_y)$ and
284: $\bm{B}_q=Q (2xz\mathbf{e}_x + 2yz\mathbf{e}_y +  (x^2+y^2-2 z^2)\mathbf{e}_z)$.
285: $\bm{B}_c$ is the constant field created by the Helmholtz coils with $B$ being the Ioffe field strength. 
286: $\bm{B}_l$ originates from the Ioffe bars and depends on the field gradient $G$.
287: $\bm{B}_q$ designates the quadratic 
288: term generated by the Helmholtz coils whose magnitude, compared to the first Helmholtz term, can be varied by changing the geometry of the trap, 
289: $Q= B\cdot \frac{3}{2}(R^2-4D^2)/(R^2+D^2)^{2}$, 
290: where $R$ is the radius of the Helmholtz coils, and $2D$ is their distance from each other.
291: 
292: If we now insert the special Ioffe-Pritchard field configuration using Eqs.~(\ref{eq:overallpotential},\ref{eq:overallfield}) 
293: into the transformed Hamiltonian (\ref{eq:UHU}) we obtain
294: \begin{align}                                                                                 \label{eq:HIP}
295:   &H_{IP} = H_A +{\bm{P}^2}/{2M} \nonumber 
296: +{BL_z}/{2}  +G(x+X)(y+Y)p_z \nonumber\\
297:   &+Q/4\big[\big((x+X)p_y-(y+Y)p_x\big)     \nonumber\\  
298:   &\cdot\big((x+X)^2+(y+Y-2(z+Z))(y+Y+2(z+Z))\big)\big]\nonumber\\
299:   &- \bm{\mu}_1 \bm B(\bm R + \bm r)-\bm{\mu}_2\bm{B}(\bm R) \; ,
300: \end{align}
301: where $H_A = {\bm p^2}/{2}-{1}/{r}$ is the operator for a field free atom.
302: The well known Zeeman term ${BL_z}/{2}$ comes from the uniform Ioffe field generated by the Helmholtz coils.
303: The following term, involving the field gradient $G$, arises from the linear field generated by the Ioffe bars and
304: couples the relative and c.m.\ dynamics.
305: The part in the squared brackets originates from the quadratic term, again created by the coils.
306: It is the only one that depends on the $Z$ coordinate, 
307: we will see below that its contribution is negligible under certain conditions. 
308: The last term couples the spin of particle two to the magnetic field. 
309: Since the electronic spins of closed shells combine to zero, the spin of particle two is the \emph{nuclear} spin only.
310: Even though $\bm \mu_2\bm B$ scales with ${1}/{M_2}$, we will still keep the term.
311: Being the only one containing the nuclear spin it is essential for a proper symmetry analysis.
312: 
313: \subsection{\label{ch:SSS}Symmetries, Scaling and the Approximation of a Single $n$-Manifold}
314: 
315: Our Hamiltonian is invariant under a number of symmetry transformations $U_S$ 
316: that are composed of the elementary operations listed in Tab.~\ref{t:sym}.
317: The parity operations $P_j$, $j \in \{x,y,z\}$, are defined by their action on the spatial laboratory coordinates of the particles 
318: which translates one-to-one to c.m. and relative coordinates.
319: In order to exchange the $x$ and $y$ components of the electronic spin we introduce the operator
320: \[S_{xy}=  \begin{pmatrix} -i&0\\ 0&1 \end{pmatrix} \; , \]
321: where $S_{xy} S_{xy}^{*}=1$. 
322: $T$ represents the conventional time reversal operator for spinless particles which, 
323: in the spatial representation,
324: corresponds to complex conjugation.
325: Our unitary symmetries are
326: \begin{subequations}\label{eq:sym}
327:   \begin{gather}
328:     P_x P_y \hat S_z \hat \Sigma_z                                    \label{eq:symI} \\
329:     P_y P_z I_{xy} S_{xy} \Sigma_{xy}                                 \label{eq:symII} \\
330:     P_x P_z I_{xy} S^{*}_{xy} \Sigma^{*}_{xy} \; .                    \label{eq:symIII}
331:   \end{gather}
332: \end{subequations}
333: The Hamiltonian is also left invariant under the antiunitary symmetry transformation
334: \begin{equation}
335: T P_y  .                                                          \label{eq:symA} \\
336: \end{equation}
337: By consecutively applying the latter operator and the unitary operators (\ref{eq:symI}), (\ref{eq:symII}) and (\ref{eq:symIII})
338: it is possible to create further antiunitary symmetries:
339: \begin{subequations}\label{eq:symAall}
340:   \begin{gather}
341:     T P_x \hat S_z \hat \Sigma_z                                        \label{eq:symAI} \\
342:     T P_z I_{xy} S_{xy} \Sigma_{xy}                                     \label{eq:symAII} \\
343:     T P_x P_y P_z I_{xy} S^{*}_{xy} \Sigma^{*}_{xy} .                   \label{eq:symAIII}
344:   \end{gather}
345: \end{subequations}
346: Paying regard to the fact that $S_{xy}^2=-\hat S_z$ and $\Sigma_{xy}^2=-\hat \Sigma_z$
347: and that $T$ neither commutes with $\hat S_y$ nor with $S_{xy}$ and $\Sigma_{xy}$,
348: one finds that the operators (\ref{eq:symI}-\ref{eq:symAIII}) form a symmetry group.
349: 
350: %%%% Table %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
351: \begin{table}[tb] 
352:   \begin{center}
353:     \begin{tabular}{lll}
354:       \hline \\[-12pt]
355:       operator &            & operation\\
356:       \hline \hline \\[-9pt]
357:       $P_{x}$ & $x$ parity & $x\rightarrow -x$, $X\rightarrow -X$ \\
358:       $\hat S_x$ & electronic spin $x$ op. & $S_y\rightarrow -S_y$, $S_z\rightarrow -S_z$ \\
359:       $\hat \Sigma_x$ & nuclear spin $x$ op. & $\Sigma_y\rightarrow -\Sigma_y$, $\Sigma_z\rightarrow -\Sigma_z$ \\
360:       $I_{xy}$ & coordinate exchange & $x\leftrightarrow y$, $X\leftrightarrow Y$ \\
361:       $S_{xy}$ & el.\ spin component exc.& $S_x\rightarrow -S_y$, $S_y\rightarrow S_x$ \\
362:       $\Sigma_{xy}$ & nuclear spin comp. exc.&  $\Sigma_x\rightarrow -\Sigma_y$, $\Sigma_y\rightarrow \Sigma_x$ \\
363:       $T$ & conventional time reversal & $A\rightarrow A^{*}$ \\
364:       \hline \hline \\[-25pt]
365:     \end{tabular}
366:   \end{center}
367:   \caption[Symmetry operation nomenclature]{\label{t:sym} Symmetry operation nomenclature. 
368:     $P_j$, $\hat S_j$, and $\hat \Sigma_j$ are exemplified by $j=x$, but hold of course also for $j=y,z$.}
369: \end{table}
370: %%%% Table %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
371: 
372: 
373: If no Ioffe field is present ($B=0$), eight additional symmetries can be found leaving the Hamiltonian invariant.
374: For an effective one particle approach (and the corresponding one particle symmetries) 
375: this was discussed in Ref.~\cite{igor:pra70:4}.
376: 
377: %% CONTINUOUS SYMMETRY %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
378: 
379: As indicated before, the quadratic magnetic field term is small 
380: and can be tuned by changing the trap geometry. 
381: It can provide a longitudinal confinement which may be treated by perturbative methods. 
382: In the case of negligible quadratic field $\bm B_q$, which we assume in the following, 
383: the term in the squared brackets of the Hamiltonian (\ref{eq:HIP}) drops out 
384: and the $Z$ coordinate is cyclic. 
385: The corresponding conjugated momentum $P_z$ is consequently conserved 
386: and the longitudinal motion is integrated by simply employing plane waves 
387: $| k_Z\rangle = \exp\{i Z k_Z\}$.
388: The constraints for this approximation to be valid can be obtained by comparing 
389: the above-mentioned term in squared brackets with the Zeeman term, $BL_z/2$.
390: Estimating $\langle x\rangle \approx n^2$,
391: $\langle xp_y\rangle \approx \langle yp_x\rangle \approx n$, and
392: using $|Q|\lessapprox B/(D^2+R^2)$
393: we find
394: \begin{align}
395:   D^2+R^2 &\gg n^4 \quad\quad  \textrm{and} \label{eq:constraintA}\\
396:   \sqrt[3]{n(D^2+R^2)} &\gg  X,Y \; ,\label{eq:constraintB}
397: \end{align}
398: where $D$ and $R$ characterize the trap geometry.
399: Eqs.~(\ref{eq:constraintA},\ref{eq:constraintB}) are easily fulfilled. 
400: We are therefore left with the Hamiltonian 
401: \begin{equation}                                                                     \label{eq:Hnotscaled}
402:   H = H_A + (P_x^2+P_y^2)/{2} + H_e \; ,
403: \end{equation}
404: where the electronic Hamiltonian reads 
405: \begin{equation}                                                                     \label{eq:Henotscaled}
406:   H_{e} = {BL_z}/{2}  +G(x+X)(y+Y)p_z - \bm{\mu}_1 \bm B(\bm R + \bm r) \; .
407: \end{equation}
408: 
409: 
410: 
411: %% $n$-mixing and scaling trafo %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
412: For all laboratory fields one finds the magnetic field strength $B$ and the magnetic field gradient $G$ 
413: to be a lot smaller than 1. 
414: Our Hamiltonian (\ref{eq:HIP}) is thus dominated by $H_A$.
415: The energies of the field free spectrum $E_A^n= -1/2 n^2$ are $n^2$-fold degenerate.
416: We can assume the Ioffe-Pritchard field not to couple adjacent $n$-manifolds 
417: as long as $|{E_A^{n}-E_A^{n\pm 1}}|/{E_{Zee}} \gg 1$. 
418: The resulting constraints $B \ll n^{-4}$,  $G\ll n^{-6}$ and $GR\ll n^{-4}$ 
419: yield $B \ll 2900$~G, $G\ll 6\cdot 10^{6}$~T/m for $n=30$ and $R\ll 2.9$~mm 
420: if we additionally assume the field gradient $G$ to be as large as $100$~T/m. 
421: In our parameter regime each $n$-manifold can therefore be considered separately.
422: We thus project the full Hamiltonian on 
423: the hydrogenic eigenfunctions $| \alpha \rangle = | n, l, m_l, m_s \rangle$, $H_A | \alpha \rangle = E_A^n | \alpha \rangle $,
424: with fixed principal quantum number $n$, that cover an entire n-manifold.
425: $l$ denotes the orbital angular momentum quantum number, 
426: $m_l$ the one of its $z$ component $L_z$
427: and $m_s$ stands for the quantum number of the electronic spin.
428: 
429: 
430: 
431: Working in a single $n$-manifold 
432: we can reformulate the term in the Hamiltonian (\ref{eq:Hnotscaled})
433: involving the field gradient $G$ into a more compact form.
434: We first consider the commutator
435: $[yz, H_A] = [yz, \bm p^2]/2 = i (yp_z + zp_y)$. 
436: This yields 
437: \begin{equation}
438:   \langle\alpha | yp_z |\alpha^{\prime}\rangle + \langle\alpha | zp_y |\alpha^{\prime}\rangle 
439:   = -i \langle\alpha | [yz,H_A] |\alpha^{\prime}\rangle = 0 \; ,
440: \end{equation}
441: since $|\alpha\rangle$ and $|\alpha^\prime\rangle$ are eigenkets 
442: to the same eigenvalue $E_n$.
443: Establishing the relation to the orbital angular momentum operator 
444: via $yp_z= L_x + zp_y$ results in
445: \begin{equation}
446:   (\langle\alpha | yp_z |\alpha^{\prime}\rangle) = \frac{1}{2} (\langle\alpha | L_x |\alpha^{\prime}\rangle) \; . 
447: \end{equation}
448: The same procedure can be applied to $xp_z$ leading to 
449: \begin{equation}
450:   (\langle\alpha | xp_z |\alpha^{\prime}\rangle) = -\frac{1}{2} (\langle\alpha | L_y |\alpha^{\prime}\rangle) \; . 
451: \end{equation}
452: Furthermore $\langle \alpha | XYp_z | \alpha' \rangle = 0$ 
453: since $p_z \sim [H_A,z]$, and eventually we can write
454: \begin{equation}
455:   G(x+X)(y+Y)p_z = G(xyp_z+XL_x/2-YL_y/2)\; ,
456: \end{equation}
457: where we omitted the bracketing alphas, 
458: but keep in mind that the above identity holds in a single $n$-manifold only. 
459: 
460: In order to remove the separate dependencies on the field parameters $B$, $G$, and on the mass $M$ from the coupling terms, 
461: we introduce scaled c.m.~coordinates, 
462: $\mathbf{R}\rightarrow \gamma^{-\frac{1}{3}} \mathbf{R}$, with $\gamma=G M$, 
463: and simultaneously we introduce the energy unit $\epsilon=\gamma^\frac{2}{3}/M$. 
464: Introducing the effective magnetic field
465: \begin{equation}
466:   \bm{G}(X,Y) = \left( \begin{array}{ccc} X \\ -Y \\ \zeta \end{array} \right) \; ,
467:   \quad \zeta =BM\gamma^{-\frac{2}{3}}\; ,
468: \end{equation}
469: and omitting the constant energy offset $E_A^n$, 
470: the Hamiltonian can be given the advantageous form
471: \begin{equation}                                                                                        \label{eq:Hwork}
472:   \mathcal H= \frac{P_x^2+P_y^2}{2}
473:   + \bm{\mu}\cdot\bm{G}(X,Y) 
474:   + \gamma^{\frac{1}{3}} (xyp_z+x S_x-y S_y).
475: \end{equation}
476: The first term is the c.m.~kinetic energy. 
477: $\bm \mu$ is the $2n^2$-dimensional matrix representation of the total magnetic moment of the electron,
478: $\frac{1}{2}(\bm L_{\bm r} +2\bm S)$, 
479: and the second term in (\ref{eq:Hwork}) describes its coupling to the effective magnetic field $\bm{G}$. 
480: The latter results from the original field~$\bm B_c+\bm B_l$ in Eq.~(\ref{eq:overallfield})
481: taking into account the corresponding coordinate and energy scaling factors. %(<-10.04.06)
482: $S_i$ are the components of the electronic spin, $\bm S=-\bm\mu_1$. 
483: The nuclear spin term $-\bm{\mu}_2\cdot\bm{B}(\bm R)$ has been omitted since it is several orders of magnitude smaller than the electronic one.
484: 
485: 
486: \section{\label{s:aa}Adiabatic Approach}
487: 
488: The large difference of the particles' masses and velocities in our two body system makes it plausible 
489: to adiabatically separate the electronic and the c.m.~motion.
490: The corresponding time scales differ substantially even for large principal quantum numbers $n$. 
491: However, due to the enormous level density in case of Rydberg atoms it is \emph{a priori} unclear 
492: whether isolated energy surfaces might exist or whether, as one might naturally assume, 
493: non-adiabatic couplings are ubiquitous and therefore an adiabatic approach might invalidate itself.
494: The procedure is reminiscent of the Born-Oppenheimer ansatz in molecular systems and is based on the idea that 
495: the slow change of the heavy particle's position  allows the electron to adapt instantaneously to the inhomogeneous field.
496: The electronic energy of the system can thus be considered as a function of the position of the heavy particle.
497: 
498: The adiabatic approximation is introduced by subtracting the transversal c.m.~kinetic energy, $\mathcal T={(P_x^2+P_y^2)}/{2}$, 
499: from the total Hamiltonian (\ref{eq:Hwork}).
500: The remaining electronic Hamiltonian for fixed center of mass reads 
501: \begin{equation}                                                                                       \label{eq:BOHe}
502:   \mathcal H_e=  \bm{\mu}\cdot\bm{G}(X,Y)   + \gamma^{\frac{1}{3}} (xyp_z+x S_x-y S_y).
503: \end{equation}
504: The electronic wave function $\varphi_\kappa$ depends parametrically on $\bm R$ 
505: and the total atomic wavefunction can be written as 
506: \begin{equation}
507:   |\Psi(\bm r,\bm R)\rangle =
508:   |\varphi_\kappa(\bm r;\bm R)\rangle
509:   \otimes |\psi_\nu(\bm R)\rangle \; ,
510: \end{equation}
511: where $|\psi_\nu(\bm R)\rangle$ is the center of mass wave function. 
512: The internal problem posed by the stationary, electronic Schr\"odinger equation
513: \begin{equation}                                                                                         \label{eq:internal}
514:   \mathcal H_e \; |\varphi_\kappa(\bm r;\bm R)\rangle = E_\kappa(X,Y)\; |\varphi_\kappa(\bm r;\bm R)\rangle
515: \end{equation}
516: is solved for the adiabatic electronic potential energy surfaces $E_\kappa(X,Y)$, 
517: that serve as a potential for the c.~m.~dynamics.
518: Within this approximation, the equation of motion for the center of mass wave function reads
519: \begin{equation}                                                                                         \label{eq:BOcm}
520:   \left( \mathcal T + E_\kappa(X,Y) \right) \; |\psi_\nu(\bm R)\rangle = \epsilon_{\nu} \; |\psi_\nu(\bm R)\rangle \; .
521: \end{equation}
522: 
523: 
524: The spatially dependent transformation $\mathcal U(X,Y)$, 
525: that diagonalizes the matrix representation $\mathcal{H}_e$ of the electronic Hamiltonian, 
526: is composed of the vector representations of the electronic eigenfunctions, 
527: $\bm{\mathcal{U}}_\kappa = \left( \mathcal{U}_{\kappa\alpha} \right) = \left( \langle \alpha |\varphi_\kappa(\bm r;\bm R)\rangle \right)$.
528: Since $\mathcal{U}$ depends on the c.m.~coordinates, the transformed kinetic energy involves 
529: non-adiabatic couplings \nolinebreak $\Delta \mathcal T$
530: \begin{equation}
531:   \mathcal{U}^\dagger \mathcal H \mathcal{U} = \mathcal{U}^\dagger \mathcal H_e \mathcal{U} + \mathcal{U}^\dagger \mathcal T \mathcal{U} = E_\kappa(X,Y) + \mathcal T + \Delta \mathcal T
532: \end{equation}
533: that have been neglected in the adiabatic approximation of Eq.~(\ref{eq:BOcm}), 
534: \begin{multline}                                                                 \label{eq:DeltaT}
535:   \Delta \mathcal T = -1/2 \cdot\big(
536:   \mathcal{U}^\dagger (\partial^2_X \mathcal{U})
537:   + \mathcal{U}^\dagger (\partial^2_Y \mathcal{U}) \\
538:   + 2 \mathcal{U}^\dagger (\partial_X \mathcal{U}) \partial_X       
539:   + 2 \mathcal{U}^\dagger (\partial_Y \mathcal{U})\partial_Y \big) \; .
540: \end{multline}
541: They can be calculated explicitly as soon as the electronic adiabatic eigenfunctions 
542: have been computed.
543: Non-adiabatic contributions can be neglected if the conditions 
544: \begin{gather}                                                                                                  \label{eq:neglectdeltaT1}
545:   \rvert \frac{\langle\varphi_{\kappa\prime} | (\partial_X \mathcal H)  |\varphi_\kappa\rangle }{E_{\kappa\prime}-E_\kappa} \rvert \ll 1  
546:   \; , \quad
547:   \rvert \frac{\langle\varphi_{\kappa\prime} | (\partial_Y \mathcal H)  |\varphi_\kappa\rangle }{E_{\kappa\prime}-E_\kappa} \rvert \ll 1 
548:   \; ,\\ \label{eq:neglectdeltaT2}
549:   \rvert \frac{\langle\varphi_{\kappa\prime} | (\partial^2_X \mathcal H)  |\varphi_\kappa\rangle }{E_{\kappa\prime}-E_\kappa} \rvert \ll 1  
550:   \; , \quad
551:   \rvert \frac{\langle\varphi_{\kappa\prime} | (\partial^2_Y \mathcal H)  |\varphi_\kappa\rangle }{E_{\kappa\prime}-E_\kappa} \rvert \ll 1 
552:   \phantom{\; ,}
553: \end{gather}
554: are fulfilled \cite{igor:pra72}. 
555: The energy denominator in (\ref{eq:neglectdeltaT1}) and (\ref{eq:neglectdeltaT2}) indicates that
556: one can expect non-adiabatic couplings to become relevant between the adiabatic energy surfaces
557:  when they come very close in energy, i.e.\ in the vicinity of avoided crossings.
558: 
559: %% Mirror symmetries of the electronic Hamiltonian %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
560: Recalling the results of the symmetry analysis, 
561: it can be demonstrated that the energy surfaces $E_\kappa$, exhibit three mirror symmetries.
562: Within the adiabatic approximation,  
563: $X$ and $Y$ are parameters in the electronic Schr\"odinger equation. 
564: Symmetry operations applied to the electronic Hamiltonian 
565: thereby merely act onto the electronic subspace.  
566: If we apply the corresponding restricted symmetry operation $U_{P} = P_x P_y \hat S_z \hat \Sigma_z$ (\ref{eq:symI}),
567: that was already shown to leave the full Ioffe-Pritchard Hamiltonian (\ref{eq:HIP}) invariant,
568: to the electronic Hamiltonian $H_e$ (\ref{eq:Henotscaled}), we find 
569: \begin{equation}
570:   U_{P}^\dagger  H_e(\bm r; X,Y)  U_{P} =  H_e(\bm r; -X,-Y) \; .
571: \end{equation}
572: Since unitarily equivalent observables, $A$ and $U^\dagger A U$,
573: possess the same eigenvalue spectrum, 
574: we find the energy surfaces to be inversion symmetric with respect to the origin in the $X$-$Y$ plane.
575: The symmetry operator $U_{Y}=T P_y$, 
576: and the operator that is composed of $U_{Y}$ and $U_{P}$, namely
577: $U_{X}=T P_x \hat S_z \hat \Sigma_z$ 
578: (see (\ref{eq:symA}) and (\ref{eq:symAI})),
579: mirror the energy surfaces at the axes, 
580: \begin{align}
581:   U_{Y}^\dagger H_e(\bm r; X,Y) U_{Y} &=  H_e(\bm r; X,-Y) \; ,\\
582:   U_{X}^\dagger H_e(\bm r; X,Y) U_{X} &=  H_e(\bm r; -X,Y) \; .
583: \end{align}
584: 
585: %% first paragraph computational approach %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
586: The electronic problem (\ref{eq:internal}), with the core fixed at an arbitrary position, is three-dimensional. 
587: No symmetry arguments can be exploited to reduce the dimensionality of the problem. 
588: In order to solve it, we employ the variational method, which maps the stationary Schr\"odinger equation
589: onto an ordinary algebraic eigenvalue problem. 
590: Since the matrix representation of the electronic Hamiltonians is sparsely occupied, 
591: an Arnoldi decomposition is used.
592: Both, this decomposition and the surfaces' mirror symmetries, help to reduce the computational cost
593: of solving the electronic Schr\"odinger equation. 
594: 
595: 
596: 
597: \section{\label{s:ees}Electronic potential energy surfaces}
598: In this section the properties of the electronic adiabatic energy surfaces are analyzed 
599: for different regimes of Ioffe field strengths and field gradients.
600: These two parameters can be used 
601: to shape the potential in which the center of mass dynamics takes place.
602: To understand how this takes place, 
603: we inspect the electronic Hamiltonian to
604: unravel the influence of the individual terms for different parameter regimes.
605: 
606: The characteristic length scale of the center of mass dynamics 
607: is of the order of one in scaled atomic units. 
608: It is therefore adequate to compare the magnitudes of the different parts of the electronic Hamiltonian (\ref{eq:BOHe})
609: in order to estimate their impact on the center of mass motion, putting $X$ and $Y$ equal to one. 
610: The first part, $\bm{\mu}\cdot\bm{G}(X,Y)$, 
611: consists of the coupling terms $X(\frac{1}{2}L_x+S_x) - Y(\frac{1}{2}L_y+S_y)$, 
612: that are then of the order of $\langle L_i \rangle \approx n$ for high angular momentum states, 
613: and of the Zeeman term $\zeta(\frac{1}{2}L_z+S_z)$, which can be as large as $\zeta n$. 
614: The second part, $\gamma^{{1}/{3}}(xyp_z+x S_x-y S_y)$, 
615: is quadratic in the relative coordinates 
616: which makes it particularly important for high principal quantum numbers $n$.
617: If we consider the expectation values of the relative coordinates to be of the order of $n^2$, and 
618: $\langle yp_z \rangle \approx \langle L_x \rangle \approx n$, 
619: the overall magnitude can be estimated to $\gamma^{{1}/{3}} n^3$.
620: In a nutshell, we have for the mentioned three terms the following relative orders of magnitude, 
621: \begin{equation}                                                                                                       \label{eq:factors}
622:   1  \; , \quad  \zeta \quad \textrm{and} \quad  \gamma^{\frac{1}{3}} n^2 \; .
623: \end{equation}
624: Due to the special form of the electronic Hamiltonian, 
625: changing the magnetic field parameters~$B$ and~$G$ while keeping their ratio~$\zeta/ \gamma^{{1}/{3}}=B/G$ (and $n$) constant 
626: results in a mere scaling of the c.m.~coordinates. 
627: We provide typical examples for values of the quantities (\ref{eq:factors}) in table \ref{t:parameters}.
628: 
629: %% table %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
630: \begin{table}[tbp]
631: \centering
632: \begin{small}
633: \begin{tabular}{lr|ccccccc|}
634: \\[-5pt]
635: \multicolumn{3}{r}{G [T/m]\quad\; \textbf{0.01$\phantom{||}$}} & \textbf{0.1} & \textbf{1} & \textbf{10} & \textbf{100} & \textbf{1000} & \multicolumn{1}{l}{\textbf{10000}} \\ 
636: \hline\\[-12pt]
637: \multicolumn{1}{r}{\textbf{}} & \multicolumn{1}{r}{n\, } & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} \\ \cline{3-9}
638: \multicolumn{1}{r}{\textbf{}} & & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} &  &\\ [-12pt]
639: {\colorbox{mygrey}{$\gamma^{\frac{1}{3}}n^2$}} & \textbf{3} & 0.001 & 0.001 & 0.003 & 0.006 & 0.014 & 0.030 & 0.064 \\ 
640:  & \textbf{10} & 0.007 & 0.015 & 0.033 & 0.071 & 0.153 & 0.329 & 0.709 \\ 
641:  & \textbf{30} & 0.064 & 0.138 & 0.296 & 0.638 & 1.375 & 2.963 & 6.383 \\ 
642:  & \textbf{50} & 0.177 & 0.382 & 0.823 & 1.773 & 3.820 & 8.229 & 17.729 \\ 
643:  & \textbf{80} & 0.454 & 0.978 & 2.107 & 4.539 & 9.778 & 21.067 & 45.387  \\ \cline{3-9}
644: \\[-10pt]
645: \multicolumn{3}{c}{B [Gauss]\;\;} & \multicolumn{1}{l}{} & \multicolumn{1}{l}{} & \multicolumn{1}{l}{} & \multicolumn{1}{l}{} & \multicolumn{1}{l}{} \\ \cline{3-9}
646: \multicolumn{1}{r}{\textbf{}} & & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} & \multicolumn{1}{l}{\textbf{}} &  &\\ [-12pt]
647: {\colorbox{mygrey}{$\zeta$}} & \textbf{0.01} & 134.0 & 28.87 & 6.220 & 1.340 & 0.289 & 0.062 & 0.013 \\ 
648:  & \textbf{0.1} & 1340 & 288.7 & 62.20 & 13.40 & 2.887 & 0.622 & 0.134 \\ 
649:  & \textbf{1} & 13402 & 2887 & 622.0 & 134.0 & 28.87 & 6.220 & 1.340 \\ 
650:  & \textbf{10} & 134015 & 28873 & 6220 & 1340 & 288.7 & 62.20 & 13.40 \\ \cline{3-9}
651: \\[-8pt]
652: \hline
653: \end{tabular}
654: \end{small}
655: \caption[Parameters]{Explicit values for $\gamma^{{1}/{3}} n^2=(GM)^{{1}/{3}} n^2$ 
656:   and $\zeta=BM^{{1}/{3}}G^{-{2}/{3}}$ for $^{87}$Rb in atomic units. 
657:   The first block lists  $\gamma^{{1}/{3}} n^2$ for different values of the field gradient $G$ and for different principal quantum numbers~$n$. 
658:   The second block lists~$\zeta$ for different field gradients and for different field strengths~$B$.
659:   % Field strengths are given in Gauss, and field gradients are given in Tesla per meter. 
660: }
661: \label{t:parameters}
662: \end{table}
663: %% table %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
664: 
665: 
666: 
667: \subsection{\label{ch:regulatingcapacity}Regulating Capacity of the Ioffe Field}
668: 
669: To understand the impact of the Ioffe field strength $B$ on the adiabatic energy surfaces, 
670: we isolate its effect by suppressing other influences.
671: This can be done by choosing a relatively low field gradient $G$ and/or a small principal quantum number $n$ (see Tab.~\ref{t:parameters}). 
672: The factor $\gamma^{\frac{1}{3}} n^2$ becomes small, and the last term in Eq.~(\ref{eq:BOHe}) will hardly provide any contribution.
673: Within this regime, that we focus on in this subsection, 
674: approximate analytical expressions for the electronic adiabatic energy surfaces can be derived. 
675: We diagonalize the approximate electronic Hamiltonian 
676: \begin{equation}                                                                                \label{eq:Hetilde}
677:   \tilde H_e= \frac{1}{2} \bm{G} \, (\bm L + 2\bm S) \; .
678: \end{equation}
679: by applying the spatially dependent unitary transformation
680: \begin{equation}                                                                                                         \label{eq:diagtrafo}
681:   U_D(X,Y)=  e^{i\phi(L_z+S_z)}  e^{i\beta(L_y+S_y)} \; , 
682: \end{equation}
683: with $\phi=\arctan\frac{Y}{X}$,
684: $\cos\beta=\gamma^{-\frac{2}{3}}M_2 B|\bm{G}(X,Y)|^{-1}$ 
685: and $\sin\beta=-\sqrt{X^2+Y^2}|\bm{G}(X,Y)|^{-1}$. 
686: This yields 
687: \begin{equation}
688:   U^\dagger_D \tilde H_e U_D = \frac{1}{2} ( L_z + 2 S_z) |\bm G(X,Y)| \; 
689: \end{equation}
690: for the transformed approximate electronic Hamiltonian. 
691: The spatially dependent transformation~$U_D$ locally rotates the magnetic moment of the electron, 
692: which includes its spin and its angular momentum, 
693: such that it is parallel to the local direction of the magnetic field. 
694: The operators~$L_z$ and~$S_z$ are not identical to the ones before having applied the transformation~(\ref{eq:diagtrafo}), 
695: they are rather related to the local quantization axis defined by the local magnetic field direction~\cite{igor:prl}.
696: 
697: 
698: The adiabatic potential surfaces evaluate to 
699: \begin{align}                                                                                                  \label{eq:Ekappasmallgradient}
700:   E_\kappa(X,Y) &= \frac{1}{2} (m_l + 2m_s) |\bm G(X,Y)| \nonumber\\
701:   &= \frac{1}{2} (m_l + 2m_s) \sqrt{X^2+Y^2+\zeta^2} \; .
702: \end{align}
703: The possible combinations of $m_l$ and $m_s$ yield $2n+1$ energy surfaces. 
704: The surfaces highest and lowest in energy correspond to circular states, ($|m_l|=l_{max}=n-1$, $m_l+2m_s=\pm n$), 
705: and they are the only non-degenerate ones. 
706: For the other surfaces ($|m_l+2m_s| < n$), 
707: the multiplicity of $(m_l+2m_s)$, 
708: and with that the degree of degeneracy of the corresponding surfaces, 
709: is given by $2n - |m_l+2m_s+1| - |m_l+2m_s-1|$. 
710: Starting from the highest energy surface, 
711: the levels of degeneracy thus are 1, 2, 4, 6, \dots. 
712: 
713: The approximate surfaces $E_\kappa$ (\ref{eq:Ekappasmallgradient}) are rotationally symmetric around the $z$-axis. 
714: An expansion around this axis ($\rho = \sqrt{X^2+Y^2} \ll \zeta$) 
715: yields a harmonic potential, 
716: \begin{equation}                                                                                                        \label{eq:Eksmallrho}
717:   E_\kappa(\rho) \approx (\zeta + \frac{1}{2\zeta} \rho^2)\cdot\frac{1}{2} (m_l + 2m_s) \; ,
718: \end{equation}
719: while we find a linear behavior, 
720: \begin{equation}                                                                                                         \label{eq:Ekbigrho}
721:   E_\kappa(\rho) \approx \frac{\rho}{2}\cdot(m_l + 2m_s) \; ,
722: \end{equation} 
723: when the center of mass is far from the $z$-axis ($\rho \gg \zeta$). 
724: 
725: % figure: n=3 different zetas
726: \begin{figure}[tbh!]
727:   \centering
728:   \includegraphics[width=8cm]{sectionx_3_different_zetas.eps}
729:   \caption[Sections for increasing Ioffe field, $n=3$.]{Sections along the $X$-axis through the electronic adiabatic energy surfaces 
730:     of an entire $n=3$ manifold. The field gradient is fixed at $G=1$ Tesla/m in order to suppress the influence 
731:     of the last term in $H_e$ (\ref{eq:BOHe}).
732:     From left to right, $\zeta=B M \gamma^{-{2}/{3}}$ increases due to an increasing Ioffe field.}                     \label{f:differentzetas}
733: \end{figure} % erstellt mit matlab: printsectionsmallbeta3
734: %%%%%%%%%%%%%%%%%%%%
735: 
736: For reasons of illustration we demonstrate the behavior of the adiabatic surfaces with increasing Ioffe field 
737: by means of a somewhat artificial example where other, previously neglected interactions might be more important. 
738: Fig.~\ref{f:differentzetas} shows sections through all the surfaces for $n=3$. 
739: This principal quantum number has been chosen in order to keep the sections simple while displaying the entire $n$-manifold. 
740: We employ $^{87}$Rb in this expository example although the electronic ground state of its outermost electron is $5$s. 
741: The sections have been calculated for the field gradient $G=1$~T/m and for different field strengths~$B$ 
742: using the total electronic Hamiltonian (\ref{eq:BOHe}). % and employing $^{87}$Rb. 
743: These parameters yield $\gamma^{{1}/{3}} n^2=0.003$, and values for~$\zeta$ ranging from $0.01$ to $1$. 
744: %
745: The surfaces in the different graphs of Fig.~\ref{f:differentzetas} indeed 
746: validate the approximate expression~(\ref{eq:Ekappasmallgradient}): 
747: We find $2n+1$ degenerate surfaces and 
748: the harmonic behavior for $|X| \ll \zeta$ gives way to a linear increase for $|X| \gg \zeta$. 
749: The energetic distances and lengths in the different graphs are comparable, 
750: since the scaling factor for the center of mass coordinates $\gamma=c_2M$ has not been changed. % for this series of plots
751: We can conclude that increasing the Ioffe field strength $B$ separates the surfaces from each other.
752: 
753: % n=30 different alphas
754:  \begin{figure}[tb!]
755:    \centering
756:      \includegraphics[width=7.5cm]{fig2sectionn30smallgamma.eps} 
757:      \caption[Sections for increasing Ioffe field, $n=30$.]{Sections along the $X$-axis through the uppermost $21$ surfaces 
758:        of the $n=30$ manifold of $^{87}$Rb for increasing ratios $B/(G n^2)$. 
759:        The field gradient is fixed at $G=10$ T/m while the Ioffe field is increased from top left to bottom right. 
760:        ($B=24$ mG, $B=48$ mG, $B=0.24$ G, $B=0.48$ G). 
761:        For small ratios $B/(G n^2)$  the influence of the second term in (\ref{eq:BOHe}) is not completely suppressed 
762:        as can be seen from the lifted degeneracies in the upper subfigures.}                      \label{f:n30differentzetas}
763:  \end{figure} % erstellt mit matlab: prav4/newpics/printsectionn30smallgamma.m
764: 
765: The data presented in Fig.~\ref{f:n30differentzetas} have been computed for the $n=30$ manifold. 
766: In order to keep the last term in (\ref{eq:BOHe}) small, the field gradient has been set to $G=0.1$~T/m 
767: ($\rightarrow \gamma^{{1}/{3}} n^2=0.14$).
768: The uppermost $21$ energy surfaces are shown for different values of the magnetic field strength $B$.
769: Similar to the $n=3$ case, one can see the harmonic behavior around the origin. 
770: The surfaces' minimal distance becomes larger for increasing $\zeta$.  
771: Since $\zeta$ and $\gamma^{{1}/{3}} n^2$ are of the same order of magnitude in subfigure (a), 
772: the contribution of the last term in (\ref{eq:BOHe}),
773: that lifts the degeneracy of the curves, is visible. 
774: 
775: 
776: %%%%%%%%%%%%%
777: The energetic distance of the approximate surfaces described by Eq.~(\ref{eq:Ekappasmallgradient}) 
778: increases with larger distances from the $Z$-axis, $\rho$, 
779: and with larger $\zeta$. 
780: The minimum energetic gap between two adjacent surfaces is at the origin and reads 
781: \begin{equation}                                                                                                            \label{eq:mindist}
782:   | E_\kappa(O) - E_{\kappa\pm 1}(O) | 
783: = \frac{B}{2} M\gamma^{-\frac{2}{3}} 
784: = \frac{\zeta}{2} \; . 
785: \end{equation}
786: The parameter $\zeta$ (an hence the field strength $B$) is the tool to control the energetic distance between the adiabatic surfaces.
787: Increasing $\zeta$, one can thus also minimize the non-adiabatic couplings $\Delta \mathcal T$ (\ref{eq:DeltaT}) 
788: discussed in Sect.~\ref{s:aa}, 
789: since they scale with the reciprocal energetic distance of the surfaces. 
790: 
791: % %% Table %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
792: %  minimal distance %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
793:    \begin{table}[bt]                                                                                         
794:    \centering
795:        %\\[-10pt]
796:      \begin{tabular}{cccccc} %{d{3}d{3}d{3}d{1}d{3}d{3}}
797:        $\textrm{B [G]}$&$\zeta$&$\textrm{G [T/m]}$&$\gamma^{1/3} n^2$&$\Delta E $&$\Delta \textrm{ [\%]}$\\
798:        \hline    \\[-10pt]
799: %       $0.001$&$0.0281$&$100$&$1.375$&$0$&$\\
800:        $0.01$&$0.288$&$100$&$1.375$&$0$&\\
801:        $0.1$&$2.89$&$100$&$1.375$&$1.291$&$15.193$\\
802:        $1$&$28.87$&$100$&$1.375$&$14.421$&$1.476$\\
803: %       $0.001$&$0.134$&$10$&$0.638$&$0$&\\
804:        $0.01$&$1.340$&$10$&$0.638$&$0.600$&$11.101$\\
805:        $0.1$&$13.40$&$10$&$0.638$&$6.694$&$0.103$\\
806:        $1$&$134.0$&$10$&$0.638$&$67.006$&$0.002$\\
807:        $10$&$1340$&$10$&$0.638$&$670.07$&$0.001$\\
808: %       $0.0001$&$0.0622$&$1$&$0.296$&$0$&\\
809: %       $0.001$&$0.622$&$1$&$0.296$&$0.278$&$11.105$\\
810:        $0.01$&$6.220$&$1$&$0.296$&$3.107$&$0.104$\\
811:        $0.1$&$62.20$&$1$&$0.296$&$31.101$&$0.001$\\
812:        $1$&$622.0$&$1$&$0.296$&$311.022$&$0.000$\\
813:        \\[-10pt]
814:        \hline
815:      \end{tabular}
816:    \caption[Minimal Distance]{Minimal distance $\Delta E$ of the two uppermost surfaces of the $n=30$ manifold.
817:      $\Delta$ denotes the discrepancy between $\Delta E$ and the approximate predicted value for the distance, $\frac{\zeta}{2}$, 
818:      according to Eq.~(\ref{eq:mindist}).} 
819:    \label{t:mindist}
820:    \end{table}
821: 
822: %%%% Table %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
823: To check the range of validity of our approximation, %area of validity, 
824: the minimal energetic distance between the two uppermost adiabatic surfaces % and the second highest surface 
825: in the $n=30$ manifold has been calculated for different parameters, 
826: subtracting the full 2D surfaces from each other, 
827: that have been obtained using the electronic Hamiltonian (\ref{eq:BOHe}). 
828: One finds the minimal distance to be located at the origin, as expected.
829: $\Delta$ in Tab.~\ref{t:mindist} denotes the relative deviation 
830: between the predicted (Eq.~(\ref{eq:mindist})) and the computed value in percent. 
831: It is small for large Ioffe field strengths $B$ and low field gradients $G$. 
832: Then we have~$\zeta \gg \gamma^{{1}/{3}}n^2$, the last term in the electronic Hamiltonian is negligible 
833: and our approximation that leads to~(\ref{eq:mindist}) is justifiable.
834: 
835: \subsection{High Gradients}                                                              \label{ch:highgradients}
836: 
837: A more complicated picture of the surface properties arises when the field gradients become larger. 
838: The last term in the electronic Hamiltonian, that accounts for finite size effects of the atom, 
839: \begin{equation}                                                 \label{eq:finitesize}
840:   \gamma^{\frac{1}{3}} (xyp_z+x S_x-y S_y) \; ,
841: \end{equation}
842: is no longer small compared to the others in equation (\ref{eq:BOHe}).
843: This results in modulations of the adiabatic surfaces we already spotted in the previous section, 
844: even though the term does not feature any dependency on $X$ and $Y$. 
845: These modulations lift the degeneracy that was found in the limit of small gradients. 
846: Their dependency on the c.m.~coordinates is introduced by the transformation $\mathcal U(X,Y)$ 
847: that diagonalizes the electronic problem (cf.~Sec.~\ref{s:aa}). 
848: 
849: % n=30, different gammas -> Kapitel {High Gradients}
850: \begin{figure}[tbp]
851:   \centering
852:   \includegraphics[width=7.5cm]{fig3sectionn30smallgamma.eps} %sectionn30_4_plusZOOMsnew.eps}
853:   \caption[Sections for increasing field gradient, $n=30$.]{Sections  ($Y=0$)
854:     through the adiabatic potential energy surfaces belonging to the 
855:     $n=30$ manifold of $^{87}$Rb for decreasing ratios $B/(G n^2)=\zeta/(\gamma^{{1}/{3}} n^2)$. 
856:     The influence of the Zeeman term in $\mathcal H_e$ (\ref{eq:BOHe}) is fixed ($\zeta=5$) 
857:     while $\gamma^{{1}/{3}} n^2$ increases. 
858:     (a) $B/(G n^2)=10 \leftrightarrow B=22.9$ mG, $G=4.81$ T/m; 
859:     (b) $B/(G n^2)=5  \leftrightarrow B=91.6$ mG, $G=38.5$ T/m; 
860:     (c) $B/(G n^2)=1  \leftrightarrow B=2.29$ G, $ G=4807$ T/m; 
861:     (d) draws the indicated region in (c) to a larger scale.}
862:   \label{f:zeta1_gammas}
863: \end{figure} % erstellt mit matlab: pra/prav4/newpics/fig3printsectionn30smallgamma
864: %in hezel@blacky:/misc/home/atph/hezel/pra/meine_parameter_pra_revise.nb
865: %fuer zeta=5,und zeta/(n^2*gamma^(1/3))=10:         B=0.022896163255175996` G=4.8074975406027`
866: %fuer zeta=5,und zeta/(n^2*gamma^(1/3))=5:          B=0.09158465302070398`  G=38.4599803248216`
867: %fuer zeta=5 und zeta/(n^2*gamma^(1/3))=1:          B=2.2896163255175983`   G=4807.4975406027`
868: 
869: In order to isolate the effect of the term (\ref{eq:finitesize}) on the adiabatic surfaces, 
870: we vary the scaling factor~$\gamma=GM$ by changing the field gradient~$G$, 
871: while keeping~$\zeta=BM^{{1}/{3}}G^{-{2}/{3}}$ constant. 
872: It is, for example, reasonable to demand $\zeta=5$ and 
873: to adjust the Ioffe field strength~$B$ to meet this condition. 
874: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
875: Fig.~\ref{f:zeta1_gammas} demonstrates the increasing influence of the interaction (\ref{eq:finitesize}) when $G$ is increased. 
876: The spectra are computed for the $n=30$ manifold of $^{87}$Rb, $\zeta=5$, 
877: while $G$ is varied from $4.8$ to $4800$~T/m. 
878: For small field gradients ((a), $B/(G n^2)=10$), 
879: the surfaces approach the shapes predicted in the limit addressed in the previous subsection (\ref{ch:regulatingcapacity}): 
880: The adiabatic surfaces with the same value of the magnetic moment $(m_l+2m_s)/2$ are approximately degenerate. 
881: The uppermost energy is the only non-degenerate one and 
882: to the corresponding eigenstate the quantum numbers $m_l=n-1$ and $m_s={1}/{2}$ can be assigned. 
883: %
884: An increasing field gradient lifts the degeneracy and groups of curves can be observed ((b), $B/(G n^2)=5$). 
885: The energetic distance between these groups stays tunable by the bias field strength, as we elucidated above (see Eq.~(\ref{eq:mindist})).
886: %
887: For even higher field gradients, the different parts of the electronic Hamiltonian are of comparable size 
888: and finite size effects substantially alter the shape of the energy surfaces ((c), (d), $B/(G n^2)=1$). 
889: Avoided level crossings appear and non-adiabatic transitions are likely to occur.
890: %
891: The uppermost energy surface, however, proves to be very robust when the field gradient is varied.
892: It is energetically well-isolated from the other adiabatic surfaces. 
893: Its distance to the surface, that is formed by the second highest eigenvalue, 
894: only decreases significantly when the ratio $B/(G n^2)$ approaches one ((c), (d)).
895: This holds true for the entire X-Y-plane. 
896: Inspecting the full uppermost surface one furthermore finds the azimuthal symmetry, 
897: that is found for large ratios $B/(G n^2)$ (see Sect.~\ref{ch:regulatingcapacity}), 
898: to be approximately conserved. 
899: 
900: 
901: % n=30 0.01G 20T  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
902:   \begin{figure}[bt!]
903:     \centering
904:     \includegraphics[width=8.5cm]{section001G20T.eps}
905:     \caption[Section $0.01$ G $20$ T/m, $n=30$.]{
906:       Section through the $n=30$ manifold for a field strength of $0.01$ Gauss and a field gradient of $20$ T/m ($^{87}$Rb).
907:       A large number of avoided crossings can be observed.
908:       The uppermost curve, however, stays isolated from the other curves.
909:       The insets show the linear behavior of the surfaces far away from the $z$-axis.
910:     }
911:     \label{f:001G20T}
912:   \end{figure} % erstellt mit matlab: printsection2
913: % n=30 0.01G 20T  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
914: 
915: 
916: Another example for the complicated structure of the adiabatic electronic energy surfaces is shown in Fig.~\ref{f:001G20T}.
917: The data are calculated for a Ioffe field strength of $0.01$~G and a field gradient of $20$~T/m. 
918: For these parameters, the contributions of all terms in the electronic Hamiltonian are of the same order of magnitude around $X=1$. 
919: One immediately notices the large number of avoided crossings between the surfaces. 
920: The uppermost curve however remains isolated from the rest of the curves. 
921: Far away from the trap center, i.e.~for large $\rho=\sqrt{X^2+Y^2}$, 
922: the coupling term in (\ref{eq:BOHe}), 
923: $X(\frac{1}{2}L_x+S_x) - Y(\frac{1}{2}L_y+S_y)$, becomes dominant. 
924: A Zeeman like splitting of the surfaces emerges, visible in the smaller graphs on the right. 
925: 
926: 
927: \subsection{Electronic Wave Functions}
928: 
929: To characterize the electronic wave function $\varphi_\kappa(\bm r;\bm R)$,
930: that corresponds to the energy eigenvalues constituting the uppermost adiabatic surface, 
931: we analyze its radial extension, angular momentum and spin.
932: % n=30, Expectation value r %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
933: \begin{figure}[tbhp]
934:   \centering
935:   % \hspace{-10pt}
936:   \includegraphics[width=8.5cm]{fig5praprint1rall.eps} 
937:   \caption[Expectation value $\langle r \rangle_\varphi$]{
938:     Expectation value $\langle r\rangle_\varphi$ of the wave functions that correspond to the uppermost electronic energy surface 
939:     for $G=100$~T/m ($n=30$, $^{87}$Rb). 
940:     $B$ is varied yielding different values for the ratio $\zeta/\gamma^{{1}/{3}}n^2=B/Gn^2$: 
941:     (a) $0.01$~G $\rightarrow B/Gn^2=0.21 $~a.u., 
942:     (b) $0.1$~G $\rightarrow B/Gn^2=2.1 $, 
943:     (c) $1$~G $\rightarrow B/Gn^2=21 $.
944:     The depicted ranges of~$X$ and~$Y$ correspond to 30 characteristic lengths
945:     of the c.m.~motion in scaled units.} 
946: %    The inset in (b) depicts the expectation value of the square angular momentum, $\langle \bm L_{\bm r}^2 \rangle$.}
947:   \label{f:rexpall}
948: \end{figure}
949: % n=30, Expectation value r %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
950: The electronic wave function depends parametrically on the c.m.~position 
951: and is, in general, distorted compared to the field free case by the external magnetic field. 
952: This is reflected in the expectation value 
953: $\langle r \rangle_e (\bm R) = \langle \varphi_\kappa(\bm r;\bm R)| r |\varphi_\kappa(\bm r;\bm R)\rangle$ 
954: which is shown in Fig.~\ref{f:rexpall} for different ratios $B/(G n^2)$. 
955: The limits of the graphs with respect to $X$ and~$Y$ correspond to thirty characteristic lengths of the c.m.~motion. 
956: %
957: While keeping $G=100$~T/m, 
958: $B$ is increased for the different plots from left to right. %, yielding increasing ratios of $B/Gn^2$. 
959: %
960: For the smallest ratio under consideration ((a), $B/Gn^2<1$), 
961: a pronounced maximum of the expectation value $\langle r \rangle_e$ can be observed at the trap center. 
962: This maximum breaks up into four maxima arranged along the diagonals when the ratio is increased ((b), $B/Gn^2>1$), 
963: while the amplitude of the spatial variation of $\langle r \rangle_e$ decreases. 
964: %
965: For an even higher value of $B$ ((c), $B/Gn^2\gg 1$), only a marginal deviation from 
966: the hydrogenic field free value for the highest possible angular momentum quantum number remains 
967: (for $n=30$ one finds \hbox{$\langle r \rangle_H(n=30,l=29)=915$}). 
968: %
969: In the region of local homogeneity, 
970: where the magnetic field does not vary significantly over the extension of the electronic cloud 
971: (i.e.~far from the $z$-axis), 
972: the expectation value approaches the field free value in all subfigures that are shown in Fig.~\ref{f:rexpall}. 
973: %
974: In accordance with the abovementioned scaling property of the electronic Hamiltonian~$\mathcal H_e$, 
975: changing the field parameters while keeping the ratio $B/Gn^2$ unaltered only modifies the scale of the c.m.~coordinates, 
976: whereas the shape of the bright regions and the energy range of the eigenvalues are not changed. 
977: 
978: % n=30, Expectation value Lr %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
979: \begin{figure}[tbhp]
980:   \centering
981:   % \hspace{-10pt}
982:   \includegraphics[width=8.5cm]{30_0.1G_100T_30xoLi_proj_Lsquare.eps} %30_0.1G_100T_30xoLi_angle_betrag.eps} % 
983:   \caption[]{
984:     Expectation values $\langle L_x\rangle$, $\langle L_y\rangle$ and $\langle L_z\rangle$ (a,b,c, respectively) 
985:     for a ratio $B/(G n^2)=2.1$ (Ioffe field $B=0.1$ G, gradient $G=100$ T/m, $^{87}$Rb, n=30). 
986:     % for the Ioffe field $B=0.1$ G and the gradient $G=100$ T/m ($^{87}$Rb, n=30). 
987:     In (d) the projection $\Pi$ of $\langle \bm {L_r}\rangle$ 
988:     onto the local magnetic field direction $\bm G$ is displayed. 
989:     It is close to the field free maximum value for the angular momentum projection, $m_{l,max}=n-1$. 
990:     Subplot (e) shows the spatial behavior of $\langle \bm L^2\rangle$.  
991:     %, whose deviation from $\Pi(\Pi+1)$ indicates an admixture of ($m_l\ne l$)-states. 
992:     % which is barely distinguishable from $\Pi(\Pi+1)$.
993:     The range of $X$ and $Y$ corresponds to 30 times the characteristic length of the c.m.\ motion. 
994:   }
995:   \label{f:Lrsmallratio}
996: \end{figure}
997: % n=30, Expectation value Lr %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
998: 
999: Let us study the angular momentum and its orientation. 
1000: It is to be expected that for dominating Ioffe field, i.e.~for very large ratios $B/(G n^2)$, 
1001: the expectation value of the angular momentum, 
1002: $\langle\bm L_{\bm r}\rangle = (\langle L_{x}\rangle,\langle L_{y}\rangle,\langle L_{z}\rangle)$, 
1003: is oriented in the Ioffe-field direction ($z$-axis). 
1004: Since the Ioffe field in any case dominates around the origin, % ($X=Y\approx 0$), 
1005: $\langle L_x \rangle$ and $\langle L_y \rangle$ are expected to vanish at $(X,Y)=(0,0)$ while  $\langle L_z \rangle$ becomes maximal. 
1006: This behavior can be observed in Fig.~\ref{f:Lrsmallratio} where $\langle L_i\rangle$ are displayed (a,b,c) 
1007: for $B=0.1$ G and $G=100$ T/m. 
1008: These parameters yield $B/(G n^2)=2.1$. 
1009: The alignment of $\langle\bm L_{\bm r}\rangle$ and the local field direction~$\bm{G}(X,Y)$ is found to be very good 
1010: in the entire $X$-$Y$-plane (the maximum angle between the two is smaller than $3.6^\circ$). 
1011: In subplot (d) we provide the spatial behavior of the projection of $\langle\bm L_{\bm r}\rangle$ onto this local field axis, 
1012: $\Pi = \langle\bm L_{\bm r}\rangle\cdot {\bm{G}(\bm R)}/{|\bm{G}(\bm R)|}$.
1013: In the local homogeneity limit, $\Pi$ approaches the maximal value for $\langle L_z\rangle$, 
1014: namely $m_{l,max}=n-1$. 
1015: %
1016: In the same manner the expectation value $\langle \bm L^2\rangle$, which is displayed in subplot (e), 
1017: converges to the maximal value, $l_{max}(l_{max}+1)=n(n-1)$.
1018: %
1019: Far from the $z$-axis, the uppermost surface hence corresponds to the circular state $|m_{l,max},l_{max}\rangle$.
1020: The deviation of $\Pi$ and $\langle \bm L^2\rangle$ 
1021: from the maximal values close to the $z$-axis reflect the admixture of states 
1022: with lower quantum numbers $m$ and $l$ to the state of the uppermost surface. 
1023: 
1024: % n=30, Expectation value Lr %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1025: \begin{figure}[tbhp]
1026:   \centering
1027:   % \hspace{-10pt}
1028:   \includegraphics[width=5.5cm]{30_1Gonlyprojection.eps}
1029:   \caption[]{
1030:     Spatial dependence of the projection $\Pi$ of  $\langle \bm {L_r}\rangle$ onto the local field axis for $B=1$ G 
1031:     (all other parameters are the same as in Fig.~\ref{f:Lrsmallratio}). 
1032:     For this ratio, $B/(G n^2)=21$, the deviations from the maximal value $m_{l,max}=n-1$ are marginal. 
1033:     (Equally, $\langle \bm {L}^2\rangle\approx l_{max}(l_{max}+1)$, not shown.) 
1034:   }
1035:   \label{f:Lrbigratio}
1036: \end{figure}
1037: % n=30, Expectation value Lr %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1038: 
1039: Increasing the applied Ioffe field by a factor of $10$ ($\rightarrow B/(G n^2)=21$), 
1040: decreases the angle between $\langle\bm L_{\bm r}\rangle$ and $\bm{G}(X,Y)$ by a factor of $10^2$, 
1041: i.e.\ a quasi perfect alignment is found. 
1042: As can be seen in Fig.~\ref{f:Lrbigratio}, the projection $\Pi$ now only deviates marginally from $m_{l,max}$. 
1043: Consequently, also $\langle \bm L^2\rangle$ exhibits only minor deviations from its maximum value in the whole $X$-$Y$-plane. 
1044: %
1045: For high ratios $B/(G n^2)$, the admixture is therefore marginal and 
1046: one can in a very good approximation assume the electronic state in the uppermost surface 
1047: to be the circular state $|m_{l,max},l_{max}\rangle$ for any c.m.\ position. 
1048: %
1049: Similar observations can be made considering the respective expectation values for the spin. 
1050: For the parameters in Fig.~\ref{f:Lrbigratio}
1051: the projection of $\langle\bm S\rangle$ onto $\bm G$ 
1052: differs less than $10^{-4}$ from $1/2$.
1053: %
1054: The expectation values of the examined electronic observables converge to the field free values for increasing ratios $B/(G n^2)$. 
1055: 
1056: Our findings indicate that the electronic structure of the atom is barely changed in the limit of large ratios $B/(G n^2)$. 
1057: The radiative lifetimes can hence be expected to differ only slightly from the field free ones~\cite{igor:pra72}. 
1058: 
1059: 
1060: \section{\label{s:cm}Quantized center of mass motion}
1061: 
1062: The energetically uppermost adiabatic electronic energy surface is the most appropriate to achieve confinement. 
1063: It does not suffer a significant deformation when the field gradient is increased 
1064: and it stays well isolated from lower surfaces for a wide range of parameters.
1065: Large energetic distances to adjacent surfaces suppress nonadiabatic couplings 
1066: (Eqs.~(\ref{eq:neglectdeltaT1}) and (\ref{eq:neglectdeltaT2})). 
1067: In order to obtain the quantized c.m.~states 
1068: we therefore solve the Schr\"odinger equation~(\ref{eq:BOcm}) for the c.m.~motion
1069: in the uppermost surface $E_{2n^2}$ 
1070: by discretizing the Hamiltonian on a grid. 
1071: %
1072: The wave function for the fully quantized state is hence composed of
1073: the eigenfunction $|\varphi_\kappa(\bm r;\bm R)\rangle$ of the electronic Hamiltonian in equation (\ref{eq:internal}), 
1074: the wave function for the center of mass motion in the $X$-$Y$ plane, $|\psi_\nu(\bm R)\rangle$,
1075: and the plain wave in $Z$ direction, 
1076: \begin{equation}
1077:   |\Psi(\bm r,\bm R)\rangle =
1078:   |\varphi_\kappa(\bm r;\bm R)\rangle
1079:   \otimes |\psi_\nu(\bm R)\rangle
1080:   \otimes | k_Z\rangle \; .
1081: \end{equation}
1082: 
1083: % cmdreier %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1084: \begin{figure}[tbhp]
1085:   \centering
1086:   % \hspace{-10pt}
1087:   \includegraphics[width=8.5cm]{cmnebeneinander.eps}
1088:   \caption[Probability densities of c.m.~states]{
1089:     Probability densities of the ground state and the first and tenth excited states
1090:     of the c.m.~motion in the uppermost adiabatic potential surface of the $n=30$ manifold of~$^{87}$Rb 
1091:     (from left to right). 
1092:     The Ioffe field strength is set to $B=0.1$ G and the field gradient is $G=10$ T/m.}
1093:   \label{f:cm}
1094: \end{figure}
1095: % cmdreier %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1096: 
1097: In Fig.~\ref{f:cm} the probability densities of the ground state and two excited states of the c.m.~motion 
1098: in the uppermost surface of the $n=30$ manifold of $^{87}$Rb are displayed. 
1099: These densities reflect the spatial symmetries of the electronic Hamiltonian $\mathcal H_e$ (\ref{eq:BOHe})
1100: and consequently those of the electronic energy surface. % $E_\kappa(X,Y)$.  %They are namely symmetric to the $X$- and $Y$-axis.
1101: They are computed for a Ioffe field strength $B=0.1$~G and a field gradient of $G=10$~T/m, 
1102: which yields $\zeta=13.4$ and $\zeta/ \gamma^{\frac{1}{3}}n^2= B/Gn^2= 21$. 
1103: According to the discussion in Sec.~\ref{ch:regulatingcapacity}, 
1104: the electronic surface then exhibits a harmonic behavior around the origin, 
1105: and the system resembles the two dimensional isotropic harmonic oscillator 
1106: in the potential  $E_h(X,Y) = (\zeta + \rho^2/2\zeta)\cdot {n}/{4}$ (cf.~Eq.~(\ref{eq:Eksmallrho}), $m_l=n-1$). 
1107: %
1108: 
1109: 
1110: The first two probability densities (from left to right) in Fig.~\ref{f:cm} 
1111: explicitely demonstrate the analogy to the harmonic oscillator. 
1112: The nodal structure of the tenth excited state is not due to a Cartesian product of 1D harmonic oscillators 
1113: but a different combination of the harmonic oscillators in the corresponding degenerate subspace. 
1114: 
1115: The energies of the c.m.~wave functions in the approximative potential $E_h(X,Y)$ read 
1116: \begin{equation}
1117:   \epsilon_{h,\nu}= (N_1+N_2+1) \ \omega \; , \quad N_1,N_2=0,1,2\dots \; , 
1118: \end{equation}
1119: where $\omega^2 = {n}/{2\zeta}$, 
1120: which are in very good agreement with the exact results in the regime where the electronic Hamiltonian (\ref{eq:finitesize}) is negligible. 
1121: Within this approximation, 
1122: the energy level spacing scales with the inverse square root of $\zeta$, %the Ioffe field strength,
1123: $\Delta \epsilon_{h,\nu} = \omega \sim {1}/{\sqrt{\zeta}}$, 
1124: whereas the energetic distance of adjacent surfaces scales linearly with~$\zeta$, see Eq.~(\ref{eq:mindist}).
1125: 
1126: 
1127: % rho %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1128: \begin{figure}[bth]
1129:   \centering
1130:     \includegraphics[width=7.5cm]{praprintrho.eps}
1131:     \caption[Expectation value $\langle \rho \rangle$.]{
1132:       Double logarithmic plot of the expectation value $\langle \rho \rangle$ 
1133:       for the c.m.~ground state (circles,~$\circ$) in the uppermost adiabatic energy surface ($n=30$, $^{87}$Rb).
1134:       The corresponding expectation values for the c.m.~wave function in a perfectly harmonic potential are depicted for comparison~({\tiny{+}}).
1135: %       We annotate that for the three data pairs at the far left of each data series, %three data pairs, whose values diverge the most from each other, 
1136: %       %smallest Ioffe field strength of every series of data pairs %most extreme parameter pairs 
1137: %       the two uppermost electronic surfaces are energetically degenerate at the origin, and therfore no confinement is expected.
1138:     }     \label{f:rho}
1139: \end{figure} % erstellt mit matlab: 
1140: % rho %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1141: 
1142: To describe the properties of the compound quantized state, we analyze the extension of the center of mass motion, 
1143: which can be measured by the expectation value
1144: \begin{equation} 
1145:   \langle\rho\rangle =
1146:   \langle\psi_\nu(\bm R)| \  
1147:   \sqrt{X^2+Y^2} \ 
1148:   |\psi_\nu(\bm R)\rangle     \; ,
1149: \end{equation}
1150: and the mean distance of the core and the electron $\langle r\rangle$. 
1151: %
1152: Fig.~\ref{f:rho} presents the radial expectation value $\langle\rho\rangle$ in Bohr radii for the c.m.~ground state 
1153: in the uppermost energy surface for different parameter sets of the magnetic field. 
1154: For comparison, the expectation value of the c.m.~state in a perfectly harmonic potential, 
1155: $\langle\rho\rangle_h = \frac{\sqrt{\pi}}{2} x_0 \sim \zeta^{{1}/{4}}$, is also depicted. 
1156: The characteristic length of the c.m.~motion is $x_0 =  1/\sqrt{\omega}  = \sqrt[4]{{2\zeta}/{n}}$. 
1157: (Due to the rescaling of the c.m.~coordinates with $\gamma^{-{1}/{3}}$ in Sec.~\ref{ch:ip} 
1158: this is of the order of $1$ for a wide range of parameter sets \{$B$, $G$, $n$\} in scaled atomic units (cf.~Tab.~\ref{t:parameters}).) 
1159: The expectation values for the real system, $\langle\rho\rangle$, 
1160: deviate from the straight line formed by $\langle\rho\rangle_h$, 
1161: as the ratio $B/G$ becomes very small. 
1162: Hence, by choosing large gradients and appropriate bias fields, 
1163: very tightly confining traps for highly excited atoms can be obtained 
1164: ($B=0.1$~G and $G=100$~T, for instance, give rise to a trap frequency of approximately $1.4$~MHz). 
1165: 
1166: The mean distance of the Rydberg electron from the core $\langle r \rangle$ 
1167: is calculated weighting that very quantity for a fixed c.m.~position $\langle r \rangle_e (X,Y)$, 
1168: with the probability density of the c.m.~wave function: 
1169: \begin{equation}
1170:   \langle r \rangle 
1171:   = \langle\psi_\nu(\bm R)| \  \langle\varphi_\kappa(\bm r;\bm R)| \ 
1172:   r \
1173:   |\varphi_\kappa(\bm r;\bm R)\rangle  \  |\psi_\nu(\bm R)\rangle     \; .
1174: \end{equation}
1175: It is depicted in Fig.~\ref{f:r_rho}, along with $\langle\rho\rangle$, versus the degree of excitation of the c.m.~motion~$\nu$. 
1176: $\langle\rho\rangle$ and $\langle r \rangle$ are of comparable size 
1177: due to the very tight confinement. 
1178: For a Ioffe field strength of $B=0.1$ G and a field gradient of $G=100$ T/m, for instance, 
1179: the ratio of $\langle \rho \rangle$ and~$\langle r \rangle$ for the ground state ($\nu=1$) is as small as 
1180: ${\langle \rho \rangle}/{\langle r \rangle} = 0.4$. 
1181: The extension of the c.m.~wave function is thus smaller than the extension of the electronic cloud.  
1182: This strongly supports the proposition that our Rydberg atoms cannot be considered as point-like particles. 
1183: 
1184: % Expectation values %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1185: \begin{figure}[tbhp]
1186:   \centering
1187: 
1188:   \includegraphics[width=8.5cm]{praprintexp_r_rho_all.eps} %xp_r_rho.eps}
1189:   \caption[Expectation values $\langle r \rangle$ and $\langle \rho \rangle$.]{
1190:     Comparison of the mean extension of the c.m.~wave function, 
1191:     $\langle\rho\rangle$, and the mean distance of the core and the electron, $\langle r \rangle$, for the $n=30$ manifold of $^{87}$Rb.
1192:     }                                                                                                                       \label{f:r_rho}
1193: \end{figure} % erstellt mit matlab: printexp_r_rho_all printed: expand; 20x20, font11
1194: % Expectation values %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1195: 
1196: The expectation value $\langle r \rangle$ for the electron remains nearly constant as the degree of excitation increases, 
1197: and it barely differs from the corresponding field free value (dashed line in Fig.~(\ref{f:r_rho})).
1198: As indicated previously, we find the electron to be in the circular state with $m_l=n-1$, 
1199: which features the smallest 
1200: mean square deviation of the nucleus-electron separation % radial uncertainty for a given $n$.  %Bethe Salpeter S. 18 
1201: $\langle r^2\rangle  - \langle r\rangle^2  = n^2(2n+1)/4$. 
1202: It is therefore possible, that the c.m.~and the electronic wave function do not even overlap. 
1203: This is indicated in the inset of the upper right plot in Fig.~\ref{f:r_rho} for $\nu=1$. %depicted
1204: 
1205: 
1206: 
1207: \section{\label{s:c}Conclusion}
1208: 
1209: 
1210: We have studied the quantum properties of ultracold Rydberg atoms in a Ioffe-Pritchard field configuration 
1211: and find trapped c.m.~quantum states to be readily achievable. 
1212: Our starting point is a two-body approach to the Rydberg atom. 
1213: Relativistic effects and deviations of the core potential from the Coulomb-potential 
1214: as well as diamagnetic interactions have not been taken into account, 
1215: which is well justified a posteriori. 
1216: Applying a spatially dependent unitary transformation 
1217: and additionally exploiting the major mass difference of the electron and the core, 
1218: we arrived at a two-particle Hamiltonian for highly excited atoms in an inhomogeneous field 
1219: where the appearance of the coupling of the relative and c.m.~dynamics is simplified substantially. 
1220: %
1221: Thenceforward we have concentrated on the special case of a Ioffe-Pritchard trap. 
1222: A symmetry analysis of the resulting Hamiltonian has been performed 
1223: revealing seven discrete unitary and anti-unitary symmetries. 
1224: Comparing the energetic contributions of the different interactions we find it legitimate 
1225: to limit our considerations to a single $n$-manifold 
1226: to solve the corresponding stationary Schr\"odinger equation. 
1227: Consequently an adiabatic approach was applied. 
1228: In the ultracold regime the Rydberg electron is much faster than the c.m.\ motion of the atom. 
1229: This justifies an adiabatic separation of the internal (relative) and the external (c.m.) dynamics. 
1230: The corresponding adiabatic electronic potential surfaces have been obtained by diagonalizing the electronic Hamiltonian matrix. 
1231: In the limit of large ratios of Ioffe field strength and field gradient, $B/(G n^2)$, 
1232: an approximate analytical expression for the adiabatic surfaces has been provided. 
1233: In this limit, the surfaces arrange equidistantly and all but the uppermost surface are degenerate. 
1234: The inter-surface distance is then proportional to the Ioffe field strength. 
1235: The structure of the electronic surfaces becomes more complex when this ratio decreases. 
1236: The shape of the uppermost surface and its energetic separation from others, however, 
1237: prove very robust with respect to changes of the field parameters. 
1238: We hence consider it the most appropriate to achieve confinement. 
1239: %
1240: Exploring the properties of the electronic wave functions 
1241: we find that the expectation values approach the field free values 
1242: when the ratio of the field strength and the field gradient,~$B/(G n^2)$, is increased. 
1243: This indicates that, despite the strong localization of the c.m., 
1244: the electronic structure of the atom is barely changed compared to the field free case. 
1245: %
1246: Examining the compound quantized states we have found a regime where the extension of the c.m.~wave function 
1247: falls below the extension of the electronic cloud, 
1248: i.e.~the c.m.~is stronger localized than the valence electron.
1249: In this regime Rydberg atoms in inhomogeneous magnetic fields can therefore not be considered as point-like particles. 
1250: %
1251: We conclude that the Ioffe-Pritchard trap provides a strong confinement for Rydberg atoms in two dimensions 
1252: that permits their trapping on a microscopic scale. 
1253: For such a one-dimensional guide, 
1254: a relatively weak longitudinal confinement along the $z$-axis could additionally be provided 
1255: for a non-Helmholtz configuration by the quadratic term. 
1256: %
1257: As a natural enrichment of the system one could study many atoms in that guide. 
1258: Challenging issues are to stabilize such a one-dimensional Rydberg gas or
1259: to answer the question if it is feasible to use the strong Rydberg-Rydberg interaction to create a chain of trapped atoms \cite{mayle:1d} 
1260: that could then serve as a tool for quantum information processing~\cite{lukin,sorensen,jaksch}
1261: making use of the state dependent atom-atom interaction. 
1262: 
1263: 
1264: \section{Acknowledgment}
1265: 
1266: Financial support by the Deutsche Forschungsgemeinschaft is gratefully acknowledged. 
1267: 
1268: \bibliography{pra}
1269: 
1270: \end{document}
1271: