1: \documentstyle[preprint,aps,graphicx]{revtex}
2: %\documentclass[preprint,aps,graphicx]{revtex4}
3:
4: \def\btt#1{{\tt$\backslash$#1}}
5:
6: \begin{document}
7: \title{Collisional dynamics of ultracold OH molecules in an electrostatic field.}
8: \author{ Alexandr V. Avdeenkov and John L. Bohn}
9: \address{JILA and Department of Physics, University of Colorado, Boulder, CO}
10: \date{\today}
11: \maketitle
12:
13: \begin{abstract}
14: Ultracold collisions of polar OH molecules are considered in the presence of
15: an electrostatic field. The field exerts a strong influence on both elastic
16: and state-changing inelastic collision rate constants, leading to
17: clear experimental signatures that should help disentangle the theory
18: of cold molecule collisions. Based on the collision
19: rates we discuss the prospects for evaporative cooling of electrostatically
20: trapped OH. We also find that the scattering properties
21: at ultralow temperatures prove to be remarkably independent of the details
22: of the short-range interaction, owing to avoided crossings
23: in the long-range adiabatic potential curves. The behavior of the
24: scattering rate constants is qualitatively understood in terms of
25: a novel set of long-range states of the [OH]$_2$ dimer.
26: \end{abstract}
27:
28:
29: \pacs{34.20.Cf, 34.50.-s, 05.30.Fk}
30:
31: \narrowtext
32:
33: \section{Introduction}
34: %\subsection{Background}
35: Polar molecules bring something entirely new to the field of
36: ultracold physics. As compared to the neutral atoms that have
37: been studied experimentally in the past, polar molecules possess
38: extremely strong, anisotropic interactions. It has been speculated
39: that dipolar interactions will lead to new properties in
40: Bose-Einstein condensates \cite{Santos,You,Goral} or degenerate
41: Fermi gases \cite{Shlyapnikov}. It has also been suggested that
42: polar molecules in optical lattices may be useful in implementing quantum
43: logic elements \cite{DeMille}.
44: On the experimental side, cold polar molecules may be produced in
45: several ways, including photoassociation of two distinct alkali species
46: \cite{Shaffer,Schloder},
47: buffer-gas cooling \cite{Doyle,Egorov},
48: or Stark slowing \cite{Meijer1,Meijer2,Meijer3}.
49:
50: Regardless of the method or production, collisions of molecules are
51: of paramount importance in describing the properties of the gas.
52: Collisions should also be interesting in their own right, as detailed
53: probes of intermolecular interactions. Several features of
54: the collisional dynamics of ground-state polar molecules, based on
55: a simplified ``toy'' model, were discussed in~\cite{bohnpolar}. This model
56: accounted for the interplay between dipole-dipole interactions, an external
57: electric field, and states of different parities. The dipole-dipole
58: interaction, which
59: scales with intermolecular separation $R$ as $1/R^3$, renders cold
60: molecule collisions completely different from cold atom collisions.
61: This is because a $1/R^3$ interaction is characterized by energy-independent
62: low energy cross sections in {\it all} partial waves, not just in s-waves
63: \cite{landau,Shakeshaft,Deb}. The relatively strong, long-range interactions imply
64: that molecules electrostatically trapped in weak-field-seeking states
65: are generally susceptible to state-changing collisions that can
66: rapidly deplete the trapped gas. The rates are in general far larger
67: than those of magnetic dipolar transitions in stretched-state
68: alkali atoms, owing largely to the
69: relative strength of electric, as opposed to magnetic, dipolar
70: interactions \cite{bohnpolar}.
71:
72: In this paper we address ultracold polar molecule collisions in a
73: more realistic model, considering in detail the OH radical. This
74: choice is motivated by the attractiveness of this molecule
75: for Stark slowing from a supersonic jet \cite{Meijer3,Ye}.
76: In particular, it has a $^2\Pi$ ground state with a small $\Lambda$-doublet
77: splitting, making it easily manipulated by modest-sized
78: electric fields.
79: A full treatment of cold collisions is somewhat hindered by
80: the fact that the OH-OH potential energy surface (PES) is
81: poorly known, although it is known to be very deep and strongly anisotropic
82: \cite{kuhn,harding}. It is not even known, for example, whether
83: OH molecules may suffer chemical reactions at ultralow temperatures.
84: As a point of reference, it was recently suggested that the reaction
85: F+H$_2$ $\rightarrow$ HF + H may proceed at appreciable rates at
86: ultralow temperatures, in spite of having a chemical barrier
87: height of 700 K \cite{Bala}.
88:
89: However, long-range dipole-dipole
90: forces strongly dominate the scattering of OH molecules in their
91: weak-field-seeking states. In this paper we will show that this arises
92: from strong avoided crossings in the long-range adiabatic potential
93: curves, which prevent the molecules from approaching close enough to
94: one another for exchange potentials to become important.
95: In this regard cooling and electrostatic trapping of
96: OH molecules can provide a wealth of information on the long-range
97: OH-OH interaction. Thus it appears possible to understand
98: a class of ultracold
99: OH-OH collisions without detailed knowledge of the short range PES.
100: This strategy would be an important stepping stone toward understanding
101: the full problem of ultracold OH collisions. Strong-field seekers,
102: by contrast, will in addition experience the short-range
103: interaction. The complete problem of exploring collisions of ultracold
104: polar molecules might therefore most efficiently proceed by a two-step analysis,
105: thus simplifying this very complicated problem.
106:
107: Accordingly we focus in this paper on the first step, namely, collisions of
108: weak-field-seeking states. After some discussion of the relevant
109: properties of OH molecules in Sec. II and their interactions in Sec. III,
110: we move on in Sec. IV to illustrate some prominent energy- and
111: field-dependent features in elastic and inelastic cross sections.
112: Mapping these features in experiments should help in unraveling
113: the long-range part of this puzzle. We also present a simplified
114: model of the long-range interaction, to help illustrate the
115: basic physics behind the behavior of the cross sections. It will
116: turn out that a new class of long-range bound states of the [OH]$_2$
117: dimer play a significant role in ultracold collisions of this
118: molecule.
119:
120: \section{OH molecule}
121:
122: The OH molecule has a fairly complicated internal structure,
123: incorporating rotation, parity, electronic spin, and nuclear spin
124: degrees of freedom, which are further confounded in
125: the presence of an electric field. We therefore begin by
126: describing the structure of this molecule and the simplifications
127: we impose to render our model tractable.
128:
129: %\subsection{General considerations}
130:
131: Molecules cooled to sub-Kelvin temperatures by Stark slowing will be
132: assumed to be in their electronic $^{2}\Pi$ ground state,
133: and $\upsilon =0 $ vibrational ground state. In this state OH is an
134: almost pure Hund's case (a) molecule, and has a dipole moment of
135: 1.68 D \cite{kuhn}.
136: Spin-orbit coupling involving the lone electronic spin
137: splits the ground state into $^{2}\Pi_{3/2}$ and $^{2}\Pi_{1/2}$ components,
138: of which $^{2}\Pi_{3/2}$ is lower in energy and is therefore the
139: state of greatest interest in ultracold collisions.
140: In our model we take into account just the lowest rotational
141: level of the corresponding ground state, $J=3/2$.
142: The energy of the first rotationally excited state with $|J=5/2,\Omega=3/2>$
143: is about 84K higher in energy \cite{coxon} and we will neglect this and
144: higher-lying states
145: in our scattering calculations. Such states will, however, contribute
146: rotational Feshbach resonances in realistic collisions.
147:
148: The isotopomer $^{16}$O$^1$H that we consider here has a nuclear spin
149: of $I=1/2$, which with a half- integer rotational quantum
150: number defines the OH molecule as a boson. Thus we should take into
151: account hyperfine structure to ensure the proper Bose symmetry.
152: We will see below that the inclusion of hyperfine structure is
153: also important in determining details of collision properties.
154: The calculations in an electric field also require knowing the Stark splitting
155: for OH molecules. Thus the Hamiltonian for the OH molecule in a field is
156: \begin{equation}
157: \label{hamone}
158: H^{OH}=H_{rot}+H_{fs}+H_{hfs}+H_{field}
159: \end{equation}
160:
161: Wave functions for the spatial degrees of freedom of the molecule
162: are constructed in the usual way. Namely, in the zero-electric-field
163: limit, eigenstates of parity $\varepsilon$ ($= \pm$) are given by the Hund's case (a)
164: representation:
165: \begin{equation}
166: \label{wfs}
167: |J M_{J} \Omega \varepsilon>=\frac{1}{\sqrt{2}} \bigl(
168: |J M_{J} \Omega > |\Lambda \Sigma> + \varepsilon |J M_{J} -\Omega > |-\Lambda -\Sigma>
169: \bigr),
170: \end {equation}
171: where the rotational part is given by a Wigner function:
172: \begin{equation}
173: |J M_{J} \Omega > =(\frac{2J+1}{8\pi^{2}})^{1/2} D^{J}_{M_{J} \Omega}(\theta,\phi,\kappa) ,
174: \end{equation}
175: and $\Omega=|\Sigma+ \Lambda|$ is the projection of the total electronic
176: angular momentum on the molecular axis.
177: The total spin of the molecule, ${\bf F}= {\bf J} +{\bf I}$, with laboratory
178: projection $M_{F}$, is then constructed by
179: \begin{equation}
180: \label{onestate}
181: |(JI)FM_{F}\Omega \varepsilon>=|\Lambda> |S \Sigma>
182: \sum_{M_{J},M_{I}} |J M_{J} \Omega \varepsilon> |IM_{I}><FM_{F}|JM_{J}I_{M_{I}}>,
183: \end{equation}
184:
185: The matrix elements for the Hamiltonian (\ref{hamone}) in this basis
186: can be found elsewhere~\cite{mizushima}. In compact form these matrix
187: elements are
188: \begin{eqnarray}
189: \label{matrix1}
190: <(JI)F \Omega M_{F} \varepsilon| H^{OH}|(J'I')F' \Omega' M_{F'} \varepsilon'>=\Bigl(
191: \delta_{\Omega,3/2} \delta_{\Omega',3/2} E_{3/2,3/2}+
192: \\
193: \nonumber
194: \delta_{\Omega,1/2} \delta_{\Omega',1/2} E_{1/2,1/2}+
195: (\delta_{\Omega,3/2} \delta_{\Omega',1/2}+
196: \delta_{\Omega,1/2} \delta_{\Omega',3/2} E_{3/2,1/2}
197: \Bigl) \times
198: \\
199: \nonumber
200: \delta_{J,J'} \delta_{F,F'} \delta_{\varepsilon,\varepsilon'}-
201: \mu {\cal E} \frac{1}{2}(1+(-1)^{J+J'}\varepsilon \varepsilon')(-1)^{F+F'+M_{F}+I-\Omega+1}
202: \times
203: \\
204: \nonumber
205: ([J][J'][F][F'])^{1/2}
206: \left( \begin{array}{ccc}
207: J & 1 & J' \\
208: -\Omega & 0 & \Omega'
209: \end{array} \right)
210: \left( \begin{array}{ccc}
211: F' & 1 & F \\
212: -M_{F'} & 0 & M_{F}
213: \end{array} \right)
214: \left\{ \begin{array}{ccc}
215: 1 & J & J' \\
216: I & F & F'
217: \end{array} \right\}.
218: \end{eqnarray}
219: In this expression $\mu$ is the molecular dipole moment,
220: ${\cal E}$ is the strength of the electric field, and
221: $E_{\Omega,\Omega^{\prime}}$ are matrix elements for
222: the fine structure
223: $H_{rot} +H_{fs}$ which can be found in~\cite{mizushima,miller}.
224: These values depend on the rotational constant, spin-orbit coupling constant,
225: hyperfine coupling constant, and $\Lambda$-doublet parameters of OH.
226: All of these constants can be found in~\cite{coxon}.
227:
228: Equation (\ref{matrix1}) shows that, strictly speaking, the only
229: good quantum number for the OH molecule is the projection of its
230: angular momentum on the laboratory axis, $M_F$. However, for our
231: present purposes it suffices to treat the quantum numbers as
232: ``almost good.'' For example, in view of the fact that OH is nearly
233: a purely Hund's case (a) molecule, the coupling between $\Omega = 1/2$
234: and $\Omega = 3/2$ states is fairly weak. We account for this interaction
235: perturbatively, by replacing the values $E_{3/2}$ and $E_{1/2}$
236: by the eigenvalues of the $2 \times 2$ matrix
237: \begin{equation}
238: \left( \begin{array}{cc}
239: E_{3/2} & E_{3/2,1/2} \\
240: E_{3/2,1/2} & E_{1/2} \\
241: \end{array} \right),
242: \end{equation}
243: keeping all other quantum numbers constant.
244:
245: Likewise, different values of the molecular spin $J$ are mixed in a field,
246: but this mixing is small in laboratory strength fields. The total spin
247: $F$ and the parity are far more strongly mixed. Accordingly,
248: in practice we transform the molecular state to a field-dressed basis
249: for performing scattering calculations:
250: \begin{eqnarray}
251: \label{ebasis}
252: |(\tilde{J}I)\tilde{F}M_{F} \Omega \tilde{\varepsilon};{\cal E} > \equiv
253: \sum_{JF\varepsilon} \alpha(JF\varepsilon) |(JI)FM_{F}\Omega \varepsilon>,
254: \end{eqnarray}
255: where $\alpha(JF\varepsilon)$ stands for eigenfunctions of the Hamiltonian~(\ref{hamone})
256: determined numerically at each value of the field. We will continue
257: to refer to molecular states by the quantum numbers $J$, $F$, and $\varepsilon$,
258: with the understanding that they are only approximately good in a field,
259: and that (\ref{ebasis}) is the appropriate molecular state.
260:
261: Figure 1 shows the Stark energies computed using all the ingredients
262: described above. In zero field the energy levels are primarily determined
263: by the $\lambda$-doublet splitting between opposite parity states,
264: whose value is $\Delta=0.056cm^{-1}$. The alternative parity states,
265: with $\varepsilon=-1$ (f states) and $\varepsilon=+1$ (e states) are
266: shown in Fig. 1(a) and Fig. 1(b), respectively. These states are further split
267: into hyperfine components with total spin $F=1$ and $F=2$. The Stark shift
268: is quadratic for fields below the critical field
269: ${\cal E}_{0}\equiv{\Delta}/{2\mu}$ ($\approx 1000 V/cm$ for OH).
270: For fields larger than ${\cal E}_{0}$ states with different parity are entirely
271: mixed and the Stark effect transforms from quadratic to linear.
272: In this case the molecular states are roughly equal linear combinations
273: of the zero-field $\varepsilon=-$ and $\varepsilon=+$ states (compare
274: Eqn. \ref{wfs}) \cite{schreel}:
275: \begin{eqnarray}
276: |J M_{J} \Omega \varepsilon=\pm 1> = \{ \begin{array}{cc}
277: | J M_{J} \mp \Omega > & M_{J} < 0 \\
278: | J M_{J} \pm \Omega > & M_{J} > 0
279: \end{array}
280: \end{eqnarray}
281:
282: \section{OH-OH Interaction}
283: %\subsection{Hamiltonian}
284: We will consider diatom-diatom scattering as two interacting
285: rigid rotors in their ground rotational states.
286: The complete Hamiltonian for the collision process can then be written
287: \begin{equation}
288: H=T_{1}+T_{2}+H^{OH}_{1}+H^{OH}_{2}+V_{s}+V_{\mu \mu}+V_{qq}+V_{disp},
289: \end{equation}
290: where $T_{i}$ and $H^{OH}_{i}$ are the translational kinetic energy
291: and internal motion of molecule $i$, including the electric field
292: as in Eqn. (\ref{hamone}). $V_{s}$ is the short- range exchange
293: interaction, $V_{\mu \mu}+V_{qq}+V_{disp}$ are the dipole- dipole,
294: quadrupole- quadrupole and dispersion long- range interactions
295: respectively. Explicit expression for the dipole-dipole ($\propto 1/R^3$)
296: and quadrupole-quadrupole ($\propto 1/R^5$) interactions are given
297: in Ref. \cite{avoird}. Matrix elements for the the dipole-quadrupole
298: interaction vanish for rigid rotor molecules in identical states
299: \cite{kuhn}, hence will not be considered here.
300:
301: The anisotropic potential between two interacting rigid-rotor molecules
302: is conveniently recast into a standard set of angular functions~\cite{avoird}:
303: \begin{eqnarray}
304: \label{potential}
305: V_{s}+V_{\mu \mu}+V_{qq}+V_{disp} \equiv
306: V(\omega_{A},\omega_{B},\omega,R)=\sum_{\Lambda} V_{\Lambda} A_{\Lambda}(\omega_{A},\omega_{B},\omega),
307: \end{eqnarray}
308: where $\Lambda \equiv (L_{A},K_{A},L_{B},K_{B},L)$ and the angular
309: functions are defined as:
310: \begin{eqnarray}
311: A_{\Lambda}(\omega_{A},\omega_{B},\omega)= \sum_{M_{A},M_{B},M}
312: \left( \begin{array}{ccc}
313: L_{A} & L_{B} & L \\
314: M_{A} & M_{B} & M
315: \end{array} \right)
316: D^{L_{A}}_{M_{A},K_{A}}(\omega_{A})
317: D^{L_{B}}_{M_{B},K_{B}}(\omega_{B})
318: C^{L}_{M}(\omega),
319: \end{eqnarray}
320: where $\omega_{A,B}=(\theta_{A,B}, \phi_{A,B})$ are the polar angles
321: of molecules A and B with respect to the lab-fixed quantization axis, and
322: ${\bf R}=(R,\omega)$ is the
323: vector between the center of mass of the molecules in the
324: laboratory fixed coordinate frame. The indices $K_A$ and $K_B$
325: denote the dependence of the interaction on the orientation of
326: the molecules about their own axes; in what follows we will ignore
327: this dependence, setting $K_A=K_B=0$. For the long-range part
328: of the interaction this approximates the quadrupole moment of OH
329: as cylindrically symmetric.
330:
331: The exchange potential $V_s$ is very complicated, consisting of
332: four singlet and four triplet surfaces \cite{harding}, and is moreover
333: poorly characterized. The most complete treatment of this surface to
334: date computes the lowest-energy potential for each value of
335: internuclear separation $R$ \cite{kuhn}. This potential finds
336: an extremely deep minimum at $R=2.7$ a.u. corresponding to chemically
337: bound hydrogen peroxide, and a second, shallower minimum at $R=6$ a.u.
338: due to hydrogen bonding forces. However, in cold
339: collisions, the scattering cross sections are so sensitive to
340: details of the short-range interaction that knowing the complete interaction
341: probably would not help anyway. More importantly, as we will see below,
342: collisions of the weak-field-seeking states are strongly dominated
343: by the long-range dipole-dipole interaction. Therefore, we will use
344: at small $R$ simply the hydrogen-bonding part of the potential
345: surface (see Fig. 13 of Ref. \cite{kuhn}), and we will treat this
346: part of the interaction as if it were isotropic.
347: Finally, we will assert that the spin states of the OH molecules
348: are in their stretched states, so that ordinary spin-exchange processes
349: will not play a role in these collisions.
350:
351: We express the Hamiltonian in a basis of projection of total
352: angular momentum,
353: \begin{eqnarray}
354: \label{wave}
355: {\cal M}= M_{F_{1}}+M_{F_{2}}+M_{l},
356: \\
357: M_{F_{i}}=M_{J_{i}}+M_{I_{i}},
358: \end{eqnarray}
359: where $M_{F_{i}}, M_{J_{i}}, M_{I_{i}}$ are the projections of the
360: full molecule spin, rotational motion and nuclear spin on the
361: laboratory axis respectively for each molecule. $M_{l}$ is the
362: projection of the partial wave quantum number on the laboratory
363: axis. In this basis the wave function for two molecules is
364: described as:
365: \begin{eqnarray}
366: \label{basis}
367: \Psi^{\cal M}=\sum_{1,2,l,M_{l}} \left\{|1> \otimes |2> \otimes |l M_{l}> \right\}^{\cal M} \times \psi^{{\cal M},1,2}(R),
368: \end{eqnarray}
369: where $\left\{...\right\}^{\cal M}$ is angular momentum part of
370: this wave function and $|i>$ is the wave function for each
371: molecule as described by Eq.(\ref{onestate}).
372:
373: Because the target and the projectile are identical bosons, we must take
374: into account the symmetry of the wave function under exchange. The
375: properly symmetrized wave function is then
376: \begin{eqnarray}
377: \label{swave}
378: \left\{|1> \otimes |2> \otimes |l M_{l}> \right\}^{s}=\frac{\left\{|1> \otimes |2> \otimes |l M_{l}> \right\}
379: +(-1)^{l} \left\{|2> \otimes |1> \otimes |l M_{l}> \right\}}{\sqrt{2(1+\delta_{12})}}
380: \end{eqnarray}
381:
382: Using the expansion of the intermolecular
383: potential~(\ref{potential}), the wave function~(\ref{wave}), and
384: taking into account the Wigner-Eckart theorem, we can present the
385: reduced angular matrix element as
386: \begin{eqnarray}
387: \label{matrix2}
388: <12lM_{l}||| A_{\Lambda}|||1'2'l'M_{l'}>=
389: \\
390: \nonumber
391: (-1)^{L_{A}+L_{B}+J_{1}+J_{1}'+J_{2}+J_{2}'+M_{F_{1}}' + M_{F_{2}}'-\Omega_{1}'-\Omega_{2}'+M_{l}-1}
392: \frac{(1+\varepsilon_{1}\varepsilon_{1}'(-1)^{L_{A}})}{2}
393: \frac{(1+\varepsilon_{2}\varepsilon_{2}'(-1)^{L_{B}})}{2}
394: \times
395: \\
396: \nonumber
397: ([l][l'][J_{1}][J_{1}'][J_{2}][J_{2}'][F_{1}][F_{1}'][F_{2}][F_{2}'])^{1/2}
398: \left( \begin{array}{ccc}
399: L_{A} & L_{B} & L \\
400: M_{F_{1}}-M_{F_{1}'} & M_{F_{2}}-M_{F_{2}'} & M_{l}-M_{l'}
401: \end{array} \right)
402: \times
403: \\
404: \nonumber
405: \left( \begin{array}{ccc}
406: J_{1}' & L_{A} & J_{1} \\
407: \Omega_{1}' & 0 & -\Omega_{1}
408: \end{array} \right)
409: \left( \begin{array}{ccc}
410: J_{2}' & L_{B} & J_{2} \\
411: \Omega_{2}' & 0 & -\Omega_{2}
412: \end{array} \right)
413: \left( \begin{array}{ccc}
414: L_{A} & F_{1} & F_{1}' \\
415: M_{F_{1}}-M_{F_{1}'} & -M_{F_{1}}& M_{F_{1}'}
416: \end{array} \right)
417: \times
418: \\
419: \nonumber
420: \left( \begin{array}{ccc}
421: L_{B} & F_{2} & F_{2}' \\
422: M_{F_{2}}-M_{F_{2}'} & -M_{F_{2}}& M_{F_{2}'}
423: \end{array} \right)
424: \left( \begin{array}{ccc}
425: l' & L & l \\
426: M_{l'} & M_{l}-M_{l'} & -M_{l}
427: \end{array} \right)
428: \left( \begin{array}{ccc}
429: l' & L & l \\
430: 0 & 0 & 0
431: \end{array} \right)
432: \times
433: \\
434: \nonumber
435: \left\{ \begin{array}{ccc}
436: L_{A} & F_{1} & F_{1} \\
437: I & J_{1'} & J_{1}
438: \end{array} \right\}
439: \left\{ \begin{array}{ccc}
440: L_{B} & F_{2} & F_{2} \\
441: I & J_{2'} & J_{2}
442: \end{array} \right\},
443: \end{eqnarray}
444: where $I=1/2$ is each molecule's nuclear spin.
445:
446: The reduced matrix elements of the angular functions
447: $A_{\Lambda}$ between symmetrized
448: basis states~(\ref{swave}) are
449: \begin{eqnarray}
450: <12lM_{l}||| A_{\Lambda}^{s}|||1'2'l'M_{l'}>=
451: \frac{<12lM_{l}||| A_{\Lambda}^{s}|||1'2'l'M_{l'}>+
452: (-1)^{l}<21lM_{l}||| A_{\Lambda}^{s}|||1'2'l'M_{l'}>}
453: {\sqrt{(1+\delta_{1,2})(1+\delta_{1',2'})}}
454: \\
455: \nonumber
456: \times
457: \frac{1+(-1)^{l+l'}}{2}
458: \end{eqnarray}
459: In practice, before each scattering calculation we transform the
460: Hamiltonian matrix from this basis into the field-dressed basis
461: defined by (\ref{ebasis}).
462: We solve the coupled- channel equations using a logarithmic
463: derivative propagator method \cite{Johnson} to
464: determine scattering matrices. Using these matrices we
465: calculate total state-to-state cross sections and rate constants.
466:
467:
468: \section{Results}
469: This paper considers the scattering problem for OH molecules
470: in an electrostatic field for cold and ultracold temperatures.
471: We are interested in particular in the highest energy weak-field-seeking
472: state of the ground rotational state,
473: $|(J,I)FM_{F},\Omega \varepsilon>=|(3/2,1/2)22,3/2,->$. This state is
474: indicated by the heavy solid line in Figure 1.
475: Since the quantum numbers $J$, $I$, and $\Omega$ are the same for
476: all the scattering processes we will consider, we will refer to this
477: state by the shorthand notation $|FM_F, \varepsilon \rangle$ $= |22,-> $.
478:
479: The main novel feature of OH- OH scattering, as
480: compared to atoms or nonpolar molecules, is the presence of
481: the long-range dipole-dipole interaction and its dependence on
482: the electrostatic field. Because these interactions strongly
483: mix different partial waves, it is essential that we include
484: more than one value of $l$. However, in the interest of emphasizing
485: the basic underlying physics, we have included only
486: the s- and d- partial waves. Sample calculations show that
487: higher partial waves change the results only slightly at the
488: energies considered. In this case, given the initial
489: state with $M_{F1} = M_{F2} = 2$, the only allowed values of
490: the total projection are ${\cal M} = 2,3,4,5,6$. Among these
491: channels only the one with ${\cal M}=4$ contains a contribution
492: from s-wave scattering, and so will deserve special attention
493: in what follows. In this case the total number of scattering channels
494: for all allowed values of ${\cal M}$, is 208.
495:
496: \subsection{Prospects for evaporative cooling}
497:
498: One of the goals of the present work is to revisit the conclusions
499: of Ref. \cite{bohnpolar} concerning the effectiveness of evaporative
500: cooling for electrostatically trapped molecules. To this end
501: Figure 2 plots the elastic and state-changing inelastic rate constants versus
502: field strength for two different collision energies, 100 $\mu K$ and 1$\mu K$.
503: Here ``elastic'' refers to collisions that do not change the internal
504: state of either molecule, while ``inel'' denotes those collisions
505: in which one or both molecules are converted into any other
506: states. These transitions are typically exothermic, leading to trap
507: heating. Not all these collisions produce untrapped states, however.
508: We find that the main contributions to the $K_{inel}$ are given
509: by processes in which quantum numbers $F$ and/or $M_{F}$ are changed by one.
510: In particular, the process $|22,- \rangle + |22,- \rangle$
511: $\rightarrow$ $|22,->+|21,->$ generally makes the
512: largest contribution to $K_{inel}$, especially at high electric field.
513:
514: At low field the rates are nearly independent of field, but begin to evolve
515: when the field approximately exceeds the critical field ${\cal E}_0=\Delta/2\mu$
516: where the Stark effect changes from quadratic to linear. Above
517: this field the rate constants exhibit oscillations as a function
518: of field. These oscillations provide an experimentally variable
519: signature of resonant collisions, meaning that mapping this field
520: dependence should help in untangling details of the long-range
521: OH-OH interaction. This is similar to the ability of magnetic-field
522: Feshbach resonances in the alkali atoms to yield detailed scattering
523: parameters \cite{Roberts,Loftus}.
524:
525: Following the example of ultracold atoms, we expect that evaporative
526: cooling can proceed when the ratio of elastic to inelastic collisions,
527: $K_{el}/K_{inel} \gg 1$. Figure 2
528: shows that this is hardly ever the case for large field values
529: ${\cal E} > {\cal E}_0$, except, perhaps, at very special field
530: values where $K_{inel}$ is at a minimum of its oscillation.
531: Since the losses are dominated by exothermic processes,
532: the ratio $K_{el}/K_{inel}$ in the threshold scattering limit
533: scales as the ratio $k_i / k_f$ of the incident and final channel
534: wave numbers, as can be seen from the Born approximation.
535: Thus at high electric fields, where the Stark splitting
536: is large (hence $k_f$ is large), the ratio may become more
537: favorable. In our calculations this apparently happens for
538: fields above $10^{4}$ V/cm.
539:
540: For fields below ${\cal E}_0 \approx 1000$ V/cm, a favorable ratio
541: of $K_{el}/K_{inel}$ is only somewhat more likely. For fields this low,
542: however, the maximum depth of an electrostatic trap is $\approx 8$mK,
543: as given by the magnitude of the Stark shift (Fig. 1). The temperature of
544: the trapped gas must therefore be well below this temperature. In the
545: example of a 100 $\mu$K gas, (Fig. 2a), the ratio $K_{el}/K_{inel}$
546: may indeed be favorable.
547: However, if the gas is cooled further, say to 1 $\mu$K (Fig. 2b),
548: this ratio becomes less favorable again. This is because of the Wigner
549: threshold laws: the exothermic rate $K_{inel}$ is energy-independent
550: at low energy, while the elastic scattering rate plummets to zero
551: as the square root of collision energy. Thus, in general, evaporative
552: cooling seems to be viable only over an extremely limited range of temperature
553: and field
554: range for the OH molecule, if at all. We therefore reiterate the message of
555: Ref. \cite{bohnpolar}, and recommend that cold OH molecules
556: be trapped by a far-off-resonance optical dipole trap, in their lowest
557: energy $|F|M_F| , \varepsilon \rangle = |11,+ \rangle$ states.
558:
559: At this point we emphasize an essential difference between evaporative
560: cooling of electrostatically trapped polar molecules and of magnetically
561: trapped paramagnetic molecules. For polar molecules the transition
562: from weak- to strong-field seeking states is {\it always} exothermic,
563: even in zero applied field. This is because the lower
564: member of a $\Lambda$-doublet is always strong-field-seeking (e.g. Fig. 1).
565: For paramagnetic molecules, by contrast, weak- and strong-field
566: seeking states can be nearly degenerate at low magnetic field values
567: (e.g., $^{17}$O$_2$ as discussed in Refs. \cite{Avdeenkov,Volpi}.
568: In the present case, OH is also paramagnetic and hence could in principle
569: be magnetically trapped. For example, the low-energy states with
570: $|FM_F \varepsilon \rangle$ $= |11, + \rangle$
571: might be suitable candidates. The influence of electric dipole
572: interactions on evaporative cooling of magnetically trapped OH has
573: yet to be explored.
574:
575:
576:
577: \subsection{Analysis of the long-range interaction}
578:
579: The general behavior of the rate constants in Fig. 2 can be
580: explained qualitatively by simplifying our model even further,
581: to a case that contains only the essential ingredients:
582: the dipole-dipole interaction, the $\Lambda$-doublet, and an electric
583: field \cite{bohnpolar}. Roughly speaking, the electric field has two effects
584: on the molecules: 1) it mixes molecular states of opposite
585: parity, thus creating induced dipole moments; and 2) the
586: resulting dipole-dipole interaction strongly couples
587: scattering channels with different partial waves, leading
588: to long-range couplings between two molecules.
589:
590: As a starting point in this analysis,
591: Figure 3 breaks down the elastic and inelastic rates into their
592: contributions from different values of the total projection
593: of angular momentum ${\cal M}$. This is done for the rates
594: calculated at an energy $E = 100$ $\mu$K, from Figure 2(a).
595: In both elastic and inelastic scattering, the rates are dominated
596: by the contribution from ${\cal M}=4$, which, it will be recalled,
597: is the only value of ${\cal M}$ that incorporates s partial waves
598: in the present model.
599: We will accordingly consider only this case in what follows.
600:
601: The model used to obtain the results in Figs 2 and 3 consists
602: of 32 channels for the block of the Hamiltonian matrix with
603: ${\cal M}=4$. To simplify the analysis of this block even further,
604: we focus on the sub-Hamiltonian with fixed quantum numbers
605: $F = M_F = 2$ for each molecule. This reduces the effective
606: Hamiltonian to six channels: there are three non-degenerate
607: thresholds $E_{\varepsilon_1 \varepsilon_2}$ corresponding
608: to different possible values of the field-dressed parity quantum number
609: $\varepsilon_i = \pm$.
610: For each of these three thresholds there are two channels, corresponding
611: at large $R$ to s- and d- partial waves.
612:
613: The simplified six-channel Hamiltonian matrix then consists of
614: $3 \times 3$ blocks $\hat{V}^{ll^{\prime}}$ parameterized
615: by partial wave quantum numbers $l$, $l^{\prime}$:
616: \begin{eqnarray}
617: \label{model}
618: \hat{H} =
619: \left( \begin{array}{cc}
620: \hat{V}_{diag}^{00} & \hat{V}_{\cal E}^{02} \\
621: \hat{V}_{\cal E}^{20} & \hat{V}_{diag}^{22}+\hat{V}_{\cal E}^{22}
622: \end{array} \right)
623: \end{eqnarray}
624: Here the diagonal components $\hat{V}_{diag}^{ll^{\prime}}$ include
625: the parity thresholds and the centrifugal interactions,
626: \begin{eqnarray}
627: \label{modeldiag}
628: \hat{V}_{diag}^{ll}=
629: \left( \begin{array}{ccc}
630: E_{--}+{\hbar^2 l(l+1) \over 2m R^2} & 0 & 0 \\
631: 0 & E_{-+}+{\hbar^2 l(l+1) \over 2m R^2} & 0 \\
632: 0 & 0 & E_{++}+{\hbar^2 l(l+1) \over 2m R^2}
633: \end{array} \right)
634: \end{eqnarray}
635: where the electric-field-dependent thresholds are given by
636: $E_{\varepsilon_{1}\varepsilon_{2}}
637: =E_{+}+E_{-}-({\varepsilon_{1}+\varepsilon_{2}}) \Delta
638: \sqrt{1+k^{2}}/2$, in terms of the dimensionless parameter
639: \begin{equation}
640: \label{kdef}
641: k \equiv {2 \mu{\cal E} \over \Delta}
642: \end{equation}
643: that relates the electric
644: field strength ${\cal E}$ to the zero-field lambda-doublet splitting
645: $\Delta = E_- - E_+$.
646: The simplified field-dependent dipole-dipole interaction term
647: $\hat{V}_{{\cal E}}^{ll^{\prime}}$ is readily parameterized in the
648: field-dressed basis as
649: \begin{eqnarray}
650: \label{dipole}
651: \hat{V}_{\cal E}^{ll^{\prime}}=
652: \left( \begin{array}{ccc}
653: k^{2} & \sqrt{2}k & 1 \\
654: \sqrt{2}k & 1-k^{2}& -\sqrt{2}k \\
655: 1 & -\sqrt{2}k & k ^{2}
656: \end{array} \right)
657: \frac{C^{ll^{\prime}}}{(1+k^{2})R^3},
658: \end{eqnarray}
659: whose coefficient $C^{ll^{\prime}}$, which is independent of both $R$
660: and the electric field, is given by
661: \begin{eqnarray}
662: C^{ll^{\prime}} =
663: -\mu^{2}
664: ([l][l'])^{1/2}
665: \left( \begin{array}{ccc}
666: l' & 2 & l \\
667: 0 & 0 & 0
668: \end{array} \right)^{2}
669: \frac{\Omega^{2} M_{F}^{2}(J(J+1)+F(F+1)-I(I+1))}{2(J(J+1)F(F+1))^{2}}
670: \end{eqnarray}
671: Notice that the
672: dipole-dipole interaction vanishes for s-waves, so that
673: $\hat{V}_{\cal E}^{00}=0$.
674:
675: Within this simplified model we will refer to the scattering
676: channels by the parities $\varepsilon_1$ and $\varepsilon_2$
677: of the two molecules, along with the partial wave
678: quantum number $l$. Thus the incident channel for weak-field-seekers will
679: be denoted $|\varepsilon_1 \varepsilon_2, l \rangle$ $=|--,0\rangle$.
680: Recall that all other quantum numbers ($J$, $I$, $\Omega$, $F$, $M_F$)
681: are assumed to have fixed values for each molecule.
682:
683: The explicit field-dependence in the coupling matrix (\ref{dipole})
684: explains qualitatively the behavior of our ultracold weak-field-seeking
685: molecules, which have incident quantum number $\varepsilon=-$.
686: For zero electric field ($k=0$), there is no direct dipole-dipole coupling between
687: identical molecules. There is, however, an off-diagonal coupling to
688: different channels with opposite parity,
689: as can be seen in the form of the Hamiltonian (\ref{model}).
690: This interaction
691: brings in the dipole-dipole coupling in second order, contributing
692: an effective dispersion-like potential $C_6^{eff}/R^6$, with a coefficient
693: \begin{equation}
694: \label{C6}
695: C_6^{eff} = \frac{(C^{20})^{2}}{2\Delta}
696: \propto \frac{{\mu}^{4}}{\Delta}
697: \end{equation}
698: for both s- and d- partial waves.
699: For the OH molecule this effective coefficient is $\approx 4 \times 10^4$
700: atomic units, far larger than for the alkali
701: atoms that are familiarly trapped. Thus, even in zero external field
702: the effective interaction strength of polar molecules is quite large.
703: This may imply the breakdown of the contact-potential approximation to describing
704: Bose-Einstein condensates of polar molecules, even when their dipoles
705: are not aligned by an external field \cite{Santos,You,Goral}. We note that the
706: quadrupole-quadrupole interaction is relatively unimportant, becoming
707: larger than this effective dispersion interaction only when
708: $R > \approx 3 \times 10^{5}$ a.u.
709:
710: When the field is switched on, the s-wave channels undergo a qualitative
711: change. Now the incident channel $|--,0 \rangle$ sees a direct coupling
712: to its d-wave counterpart $|--,2\rangle$, via the matrix element
713: $V^{l=0,l'=2} \propto \frac{k^{2}}{1+k^{2}} \frac{\mu^{2}}{R^{3}}$.
714: This perturbation generates an far stronger effective long-range potential
715: of the form $C_4^{eff}/R^4$, with
716: \begin{eqnarray}
717: \label{slong}
718: C_4^{eff}= -(\frac{k^{2}}{1+k^{2}})^{2} \frac{\mu^{4}2m}{l(l+1)},
719: \end{eqnarray}
720: where $l=2$. Thus the electric field is able to completely alter
721: the character of the intermolecular interaction.
722:
723: For d-wave collisions, the dipole-dipole coupling is direct, but not in
724: the limit of zero field, where the molecules are in parity eigenstates.
725: At low fields ($k \ll 1$, where the Stark effect is quadratic),
726: the diagonal coupling $V_{\cal E}^{22} \propto k^2/(1+k^2)$, is small.
727: In this limit the molecules are nearly in parity eigenstates,
728: hence do not ``know'' that they have dipole moments. At larger fields
729: this interaction grows in scale, thus ``activating'' the dipoles.
730: This is why the rate constants shown in Figs 2,3 begin to evolve
731: at fields near ${\cal E}={\cal E}_0$. It is also why the contributions
732: from all angular momentum projections except the one with ${\cal M}=0$
733: contribute only weakly to scattering at low field. The
734: channels with ${\cal M} \ne 4$ are all of d-wave character, hence
735: obey a threshold law $\sigma \propto E^{2}$ at low fields, then
736: evolve to a $\sigma \propto {\rm const}$ threshold law at larger fields.
737:
738: \subsection{Large-field oscillations and long-range states of the
739: [OH]$_2$ dimer}
740:
741: At fields larger than the critical field ${\cal E}_0$, the rate
742: constants in Fig.2 exhibit oscillations with field. Significantly, these occur
743: only when the projection of total angular momentum ${\cal M}=4$,
744: which is the only case in which s- and d- partial waves are mixed (Fig. 3).
745: To understand this oscillating behavior of cross sections we
746: show in Fig. 4(a) the adiabatic potential curves in the simplified 6-channel
747: model~(\ref{model}). In the case shown the field is ${\cal E}=10^4$ V/cm.
748: In this figure a strong avoided crossing can be seen at $R \approx 60$ a.u.,
749: corresponding to the crossing of the attractive $|\varepsilon_1
750: \varepsilon_2,l \rangle=$
751: $|--,0\rangle$ channel
752: with the repulsive $|-+,2\rangle$ channel. The strong dipole-dipole
753: interaction between these different partial waves creates the
754: adiabatic potential shown as a heavy black line and labeled $U_0$.
755:
756: This potential curve supports bound states of the [OH]$_2$ dimer.
757: These bound states are of purely long-range character, similar to
758: the long-range states of the alkali dimers \cite{Movre} that have been used
759: in photoassociation spectroscopy studies of ultracold collisions
760: \cite{Jones,Stwalley}. Moreover, in the case of the [OH]$_2$
761: states the shape of the potential $U_0$, hence the energies of the bound states,
762: are strongly subject to the strength of the applied electric field.
763: The curve in Fig. 4(a) in fact possesses no bound states in zero
764: field, but five by the time the field reaches $10^4$ V/cm.
765: More realistic adiabatic potentials are of course more elaborate,
766: as shown in Fig. 4(b) for the more
767: complete Hamiltonian that includes hyperfine levels. Nevertheless,
768: in this figure, too, can be seen adiabatic potential wells
769: that will support bound states.
770:
771: The significance of these curves is twofold: the crossing is
772: very adiabatic, implying that coupling to lower-energy channels
773: is weak, and that therefore the cross sections depend only weakly
774: on details of the short-range potentials. This we have indeed
775: verified by altering the short-range potential in the full calculation.
776:
777: Additionally, as the field strength grows and the potential becomes
778: deeper, new bound states are added to the potential, causing
779: scattering resonances to appear. This is the cause of
780: the oscillations observed in the rate constants in Fig. 2. To
781: illustrate this, we reproduce in Fig. 5 the complete elastic
782: scattering rate constant (solid line), along with the
783: same quantity as computed in the simple six-channel model
784: (dashed line). The qualitative behavior is nearly the same, namely,
785: oscillations appear at fields above ${\cal E}_0$. Moreover,
786: the arrows in the figure indicate the values of the field for which
787: a bound state of $U_0$ coincides with the scattering threshold.
788: These fields correspond fairly well
789: to the peaks, although they are somewhat offset by coupling to
790: other channels. Nevertheless, this simple picture
791: clearly identifies the origin of these oscillations with the
792: existence of long-range bound states.
793:
794: These resonant states are not Feshbach resonances, since there is
795: no excitation of internal states of the molecules; nor are they
796: shape resonances in the usual sense, since there is no barrier
797: through which the wave function must tunnel. Instead, they are the
798: direct result of altering the interaction potential to place a
799: bound state exactly at threshold \cite{landau}.
800: Probing these states through direct scattering of weak-field-seeking
801: states should reveal details about the long-range OH-OH interaction,
802: making possible a comparison of theory and experiment without
803: the need to fully understand the short-range [OH]$_2$ potential energy
804: surface.
805:
806:
807:
808: \section{Conclusion}
809: In this paper we theoretically investigated ultracold collisions of
810: ground state polar diatomic molecules in an electrostatic field,
811: taking OH molecules as a prototype. Focusing on weak-field-seeking
812: states, we have strengthened the suppositions in Ref. \cite{bohnpolar}
813: that long-range dipolar interactions drive inelastic scattering processes
814: that are generally unfavorable for evaporative cooling of this species.
815: However, at electric fields above a characteristic value ${\cal E}_0$,
816: oscillations occur in both elastic and inelastic collision rates,
817: implying that a regime may be found where the ratio
818: $K_{elastic}/K_{inelastic}$ is favorable for cooling.
819: Even though evaporative cooling may be difficult, the inelastic rates
820: may nevertheless prove useful diagnostic tools for cold collisions
821: of these molecules. The Stark slowing technique provides
822: a means of launching a bunch of molecules toward a stationary trapped
823: target, i.e., of making a real scattering experiment \cite{Meijer3,Ye}.
824:
825: For actual trapping and cooling purposes, for instance as a means of
826: producing molecular Bose-Einstein condensates or degenerate Fermi gases, it
827: seems likely that the molecules must be trapped in their strong-field
828: seeking states. Collisions of these species will present their own
829: difficulties, since they will depend strongly on the short-range
830: part of the potential energy surface. However, the scattering length
831: for OH-OH scattering may be determined by photoassociation spectroscopy
832: to the long-range bound states we have described above. This will be
833: analogous to the determination of alkali scattering lengths
834: \cite{Jones,Stwalley}, except that microwave, rather than optical, photons
835: will be used. The detailed properties of the long-range [OH]$_2$ states,
836: and prospects for using them in this way, therefore deserve further attention.
837:
838: %{\bf \Large Acknowledgment}
839: This work was supported by the National Science Foundation. We acknowledge
840: useful discussions with J. Hutson and G. Shlyapnikov.
841:
842: \begin{references}
843:
844: \bibitem{Santos} L. Santos, G.V. Shlyapnikov, P. Zoller and M. Levenstein,
845: Phys. Rev. Lett. {\bf 85}, 1791 (2000).;\\ Stefano Giovanazzi et.al., J.Phys.B: At.Mol.Opt.Phys. {\bf 34}, 4757 (2001)
846:
847: \bibitem{You} S. Yi and L. You, Phys. Rev. A {\bf 63}, 053607 (2001).
848:
849: \bibitem{Goral} K. G\'{o}ral, K. Rz\c{a}\.{z}ewski, and T. Pfau, Phys.
850: Rev. A {\bf 61}, 051601 (2000); J.-P. Martikainen, M. Mackie, and
851: K.-A. Suominen, Phys. Rev. A {\bf 64}, 037601 (2001).
852:
853: \bibitem{Shlyapnikov} M. A. Baranov, M. S. Mar'enko, V. S. Rychkov, and
854: G. V. Shlyapnikov, cond-mat 0109437 (2001).
855:
856: \bibitem{DeMille} D. DeMille, quant-ph/0109083 (2001).
857:
858: \bibitem{Shaffer} J. P. Shaffer, W. Chalupczak, and N. P. Bigelow,
859: Phys. Rev. Lett. {\bf 82}, 1124 (1999).
860:
861: \bibitem{Schloder} U. Schl\"{o}der, C. Silber, and Z. Zimmerman, Appl. Phys.
862: B {\bf 73}, 801 (2001).
863:
864: \bibitem{Doyle} J. D. Weinstein, R. deCarvalho, T. Guillet, B.
865: Friedrich, and J. M. Doyle, Nature {\bf 395}, 148 (1998).
866:
867: \bibitem{Egorov} D. Egorov,
868: J. D. Weinstein, D. Patterson, B. Friedrich, and J. M. Doyle,
869: Phys. Rev. A {\bf 63}, 030501 (2001).
870:
871: \bibitem{Meijer1} H. L. Bethlem, G. Berden, A. J. van Roij, F. M. H.
872: Crompvoets, and G. Meijer, Phys. Rev. Lett. {\bf 84}, 5744 (2000).
873:
874: \bibitem{Meijer2} H. L. Bethlem, G. Berden, F. M. H. Crompvoets,
875: R. T. Jongma, A. J. A. van Roij, and G. Meijer, Nature {\bf 406},
876: 491 (2000).
877:
878: \bibitem{Meijer3} H. L. Bethlem, F. M. H. Crompvoets, R. T. Jongma,
879: S. Y. T. van der Meerakker, and G. Meijer, to appear in Phys. Rev. A.
880:
881: \bibitem{bohnpolar} John L. Bohn, Phys.Rev.A {\bf 63}, 052714(2001)
882:
883: \bibitem{landau} L.D. Landau and E.M. Lifshitz, Quantum Mechanics: Non- Relativistic Theory
884: (Nauka, Moscow, 1989, 4th ed.)
885:
886: \bibitem{Shakeshaft} R. Shakeshaft, J. Phys. B {\bf 5}, L115 (1972).
887:
888: \bibitem{Deb} B. Deb and L. You, Phys. Rev. A {\bf 64}, 022717 (2001).
889:
890: \bibitem{Ye} J. Ye, private communication.
891:
892: \bibitem{coxon} J.A. Coxon, Can.J.Phys {\bf 58}, 933 (1980).;\\
893: Thomas D. Varberg and Kenneth M. Evenson, J.Mol.Spec. {\bf 157}, 55 (1993).
894:
895: \bibitem{kuhn} Bernd Kuhn et.al., J.Chem.Phys. {\bf 111} 2565 (1998).
896:
897: \bibitem{harding} Lawrence B. Harding, J.Chem.Phys. {\bf 95} 8653 (1991).
898:
899: \bibitem{Bala} N. Balakrishnan and A. Dalgarno, Chem. Phys. Lett.
900: {\bf 341}, 652 (2001).
901:
902: \bibitem{Johnson} B. R. Johnson, J. Comput. Phys. {\bf 13}, 445 (1973).
903:
904: \bibitem{mizushima} M.Mizushima, The Theory of Rotating Diatomic Molecules (Wiley, New York, 1975);\\
905: Helene Lefebvre- Brion, Robert W. Field, Perturbations in the Spectra of Diatomic Molecules
906: (Academic Press Inc.(London) Ltd., 1986);\\
907: Karl F. Freed, J.Chem.Phys. {\bf 45} 4214 (1966).
908:
909: \bibitem{miller} Steven M. Miller and David C. Clary, J.Chem.Phys. {\bf 98} 1843 (1993).
910:
911: \bibitem{schreel} K.Schreel and J.J.Meulen, J.Phys.Chem.A {\bf 101}, 7639 (1997).
912:
913: \bibitem{avoird} Ad van der Avoird, Topics in Curr. Chem, {\bf 93} 1 (1980)
914:
915: \bibitem{Roberts} J. L. Roberts, J. P. Burke, S. L. Cornish, N. R. Claussen,
916: E. A. Donley, and C. E. Wieman, Phys. Rev. A {\bf 64}, 024702 (2001).
917:
918: \bibitem{Loftus} T. Loftus, C. A. Regal, C. Ticknor, J. L. Bohn, and D. S. Jin,
919: submitted to Phys. Rev. Lett.
920:
921: \bibitem{Avdeenkov} Alexandr V. Avdeenkov and John L. Bohn, Phys.Rev.A {\bf 64}, 052703(2001)
922:
923: \bibitem{Volpi} A. Volpi and J. L. Bohn, to appear in Phys. Rev. A.
924:
925: \bibitem{Movre} M. Movre and G. Pichler, J. Phys. B {\bf 10}, 2631 (1977);
926: W.C. Stwalley, Y.-H. Uang, and G. Pichler, Phys. Rev. Lett.
927: {\bf 41}, 1164 (1978).
928:
929: \bibitem{Jones} K. M. Jones, P. S. Julienne, P. D. Lett, W. D. Phillips,
930: E. Tiesinga, and C. J. Williams, Europhys. Lett. {\bf 35}, 85 (1996).
931:
932: \bibitem{Stwalley} J. P. Burke, Jr. {\it et al}, Phys. Rev. A {\bf 60},
933: 4417 (1999); C. J. Williams {\it et al.}, Phys. Rev. A {\bf 60}, 4427 (1999).
934:
935: \end{references}
936:
937: \newpage
938: \begin{figure}
939: \label{stark}
940: \centerline{\includegraphics[width=0.50\linewidth,height=0.60\linewidth,angle=-90]{fig1a.ps}}
941: \centerline{\includegraphics[width=.50\linewidth,height=0.60\linewidth,angle=-90]{fig1b.ps}}
942: \caption{
943: The Stark effect in ground state OH molecules, taking into account hyperfine splitting.
944: (a) shows the states that have odd parity $\varepsilon=-$ in zero electric field
945: (f states), whereas (b) shows those of even parity (e states).
946: The weak-field-seeking state with quantum numbers $F=M_F=2$, the subject
947: of this paper, is indicated by the heavy solid line. Note that states with
948: $M_F = \pm |M_F|$ are degenerate in an electric field.
949: }
950: \end{figure}
951:
952: \begin{figure}
953: \label{rates}
954: \centerline{\includegraphics[width=0.50\linewidth,height=0.60\linewidth,angle=-90]{fig2a.ps}}
955: \centerline{\includegraphics[width=.50\linewidth,height=0.60\linewidth,angle=-90]{fig2b.ps}}
956: \caption{
957: Rate constants versus electric field for OH-OH collisions with molecules
958: initially in their $|FM_F,\varepsilon \rangle$ $=|22->$ state.
959: Shown are the collision energies $100\mu K$ (a) and $1\mu K$(b).
960: Solid lines denote elastic scattering rates, while dashed lines
961: denote rates for inelastic collisions, in which one or both molecules
962: changes its internal state.
963: These rate constants exhibit characteristic oscillations in field when
964: the field exceeds a critical field ${\cal E}_0 \approx 1000$ V/cm.
965: }
966: \end{figure}
967:
968: \begin{figure}
969: \label{elastic}
970: \centerline{\includegraphics[width=0.50\linewidth,height=0.60\linewidth,angle=-90]{fig3a.ps}}
971: \centerline{\includegraphics[width=0.50\linewidth,height=0.60\linewidth,angle=-90]{fig3b.ps}}
972: \caption{
973: Elastic (a) and inelastic (b) rate constants versus electric field for the
974: same circumstances as in Figure 2(a). The rates are separated into
975: contributions from different values of ${\cal M}$, the projection of
976: total angular momentum on the laboratory $z$-axis.
977: }
978: \end{figure}
979:
980:
981: \begin{figure}
982: \label{adiabatic}
983: \centerline{\includegraphics[width=0.50\linewidth,height=0.60\linewidth,angle=-90]{fig4a.ps}}
984: \centerline{\includegraphics[width=0.50\linewidth,height=0.60\linewidth,angle=-90]{fig4b.ps}}
985: \caption{
986: Adiabatic potential energy curves. The curves in (a) correspond to the
987: simplified six-channel model described in the text, and show a long-range
988: potential well (labeled $U_0$) that can hold bound states of the [OH]$_2$
989: dimer. The curves in (b) are those for the more complete calculation that
990: includes hyperfine structure.
991: }
992: \end{figure}
993:
994:
995: \begin{figure}
996: \label{bound}
997: \centerline{\includegraphics[width=0.50\linewidth,height=0.60\linewidth,angle=-90]{fig5.ps}}
998: \caption{
999: Elastic rate constants versus field, as in Figure 2.
1000: The solid line reproduces the elastic rate constant from figure 2. The
1001: dashed line is an approximate elastic rate constant based on the simplified
1002: six-channel model described in the text. The arrows indicate values of
1003: the electric field at which bound states of the long-range potential
1004: $U_0$ (Figure 4) coincide with the scattering threshold.
1005: }
1006: \end{figure}
1007:
1008: \end{document}
1009: