1: \documentclass[prl,twocolumn,amsmath,amssymb,groupedaddress]{revtex4}
2:
3: %\documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb,aps]{revtex4}
4: %\documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb,aps,eqsecnum]{revtex4}
5: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
6:
7: % Some other (several out of many) possibilities
8: %\documentclass[preprint,aps]{revtex4}
9: %\documentclass[preprint,aps,draft]{revtex4}
10: %\documentclass[prb]{revtex4}% Physical Review B
11:
12: %\usepackage{bm}% bold math
13: %\usepackage{dcolumn}% Align table columns on decimal point
14:
15: \newcommand{\s}{{\sigma}}
16: \newcommand{\rv}{{\bf r}}
17: \newcommand{\xv}{{\bf x}}
18: \newcommand{\kv}{{\bf k}}
19: \newcommand{\eps}{{\varepsilon}}
20: \newcommand{\uv}{{\bf u}}
21: \newcommand{\vv}{{\bf v}}
22: \newcommand{\qv}{{\bf q}}
23: \newcommand{\nv}{{\bf n}}
24: \def\mm#1{\underline{\underline{{#1}}}}
25: \newcommand{\um}{{\mm{u} }}
26: \newcommand{\Lm}{{\mm{\Lambda} }}
27: \newcommand{\pv}{{\bf p}}
28: \newcommand{\Tr}{{\rm Tr}}
29: \newcommand{\px}{{\partial_x}}
30: \newcommand{\py}{{\partial_y}}
31: \newcommand{\ppi}{{\partial_i}}
32: \newcommand{\ppj}{{\partial_j}}
33: \newcommand{\pz}{{\partial_z}}
34: \newcommand{\ppk}{{\partial_k}}
35: \newcommand{\ppl}{{\partial_l}}
36: \usepackage[dvips]{graphicx}
37:
38: \begin{document}
39: \title[]{Superfluid transitions in bosonic atom-molecule mixtures near
40: Feshbach resonance}
41: %Spontaneously
42: %Orientationally-Ordered Solids}
43: \author{Leo Radzihovsky}
44: \author{Jae Park}
45: %\email[]{radzihov@colorado.edu}
46: %\homepage[]{Your web page}
47: %\thanks{}
48: %\altaffiliation{}
49: \affiliation{Department of Physics, University of Colorado,
50: Boulder, CO 80309}
51: \author{Peter B. Weichman}
52: \affiliation{ALPHATECH, Inc., 6 New England Executive Place,
53: Burlington, MA 01803}
54:
55: %Collaboration name if desired (requires use of superscriptaddress
56: %option in \documentclass). \noaffiliation is required (may also be
57: %used with the \author command).
58: %\collaboration{}
59: %\noaffiliation
60:
61: \date{\today}
62:
63: \begin{abstract}
64: We study bosonic atoms near a Feshbach resonance, and predict that
65: in addition to a standard normal and atomic superfluid phases,
66: this system generically exhibits a distinct phase of matter: a
67: molecular superfluid, where molecules are superfluid while atoms
68: are not. We explore zero- and finite-temperature properties of the
69: molecular superfluid (a bosonic, strong-coupling analog of a BCS
70: superconductor), and study quantum and classical phase transitions
71: between the normal, molecular superfluid and atomic superfluid
72: states.
73: \end{abstract}
74: \pacs{}
75:
76: \maketitle
77:
78: Experimental realizations and coherent manipulation of
79: trapped degenerate gases \cite{JILAbec,MITbec} is leading to
80: exciting possibilities for studies of quantum liquids in
81: previously unexplored (e.g., extremely coherent and
82: nonequilibrium) regimes. Magnetic field-induced Feshbach
83: resonance (FBR) in ultracold atom collisions allows fine
84: tuning of interactions in these quantum fluids, and was recently
85: used to create a degenerate mixture of coherently-coupled alkali
86: atoms and their diatomic molecules \cite{donley}.
87:
88: \begin{figure}[bth]
89: \centering
90: \setlength{\unitlength}{1mm}
91: \begin{picture}(70,65)(0,0)
92: \put(-30,-62){\begin{picture}(60,60)(0,0)
93: \special{psfile=phase_diagramA.eps vscale=62 hscale=62}
94: \end{picture}}
95: \end{picture}
96: \caption{Mean-field phase diagram for a bosonic atom-molecule mixture,
97: showing molecular (MSF) and atomic (ASF) superfluid phases.}
98: \label{phasediagramA}
99: \end{figure}
100:
101: In this Letter we study phases and phase transitions that take
102: place in bosonic atom-molecule mixtures. Our main contribution is the
103: prediction of a thermodynamically distinct ``molecular
104: superfluid'' (MSF) phase, that, as illustrated in
105: Figs.\ \ref{phasediagramA}, \ref{phasediagramB} ubiquitously
106: intervenes between the ``normal'' (N) and ``atomic superfluid''
107: (ASF) phases. Molecular superfluidity [and accompanying
108: off-diagonal long-range molecular order (ODLRO)] distinguishes MSF
109: from the normal state, and the {\em absence} of atomic
110: superfluidity from the ASF, in which both bosonic atoms and
111: molecules display ODLRO. If atomic and molecular components can
112: be imaged independently \cite{jinBCS}, in a harmonic trap MSF
113: should be easily identifiable by a sharp Bose-Einstein
114: condensation (BEC) peak in the molecular density profile and a
115: broad, seemingly normal, thermal atomic cloud.
116:
117: As a conventional superfluid, MSF is characterized by a
118: (molecular) acoustic second-sound mode. However, MSF also exhibits
119: a {\em gapped}, Bogoliubov-like mode, derived from unpaired atom
120: excitations. MSF ground state (bosonic analog of the BCS state)
121: exhibits strong (atom and molecule) pairing correlations that in a
122: trap should be observable in the atomic density-density
123: correlation function. Experimentally, MSF should be accessible by
124: tuning temperature, atomic density (or number), and detuning
125: $\nu$. The MSF-ASF transition is in the $(d+1)$- and
126: $d$-dimensional Ising universality classes for $T=0$ and finite
127: $T$, respectively, and is {\em reentrant} as a function of
128: detuning $\nu$ and density $n$. The tricritical point, where N,
129: MSF and ASF meet, exhibits nontrivial and, to our knowledge,
130: unexplored quantum critical behavior for $d < 4$. We now sketch
131: derivation of these results.
132:
133: Near a FBR a bosonic atom-molecule system is characterized by the
134: grand-canonical Hamiltonian $\hat{H}_\mu = \hat{H} - \mu \hat{N}$
135: \cite{foot1}
136: \begin{eqnarray}
137: \hat{H}_\mu &=& \int d^d{\bf x} \Bigg[\sum_{\s=1}^2
138: \left(\hat{\psi}^\dagger_\s \hat{h}_\s \hat{\psi}_\s
139: + \frac{g_\s}{2} \hat{\psi}^\dagger_\s
140: \hat{\psi}^\dagger_\s \hat{\psi}_\s \hat{\psi}_\s \right)
141: \nonumber \\
142: &&+\ g_{12} \hat{\psi}^\dagger_1 \hat{\psi}^\dagger_2
143: \hat{\psi}_2 \hat{\psi}_1
144: - \alpha \left(\hat{\psi}^\dagger_1 \hat{\psi}^\dagger_1
145: \hat{\psi}_2 + {\rm h.c.} \right) \Bigg]
146: \label{hamiltonian}
147: \end{eqnarray}
148: where $\hat{\psi}^\dagger_\s({\bf x}),\hat{\psi}_\s({\bf x})$ are
149: bosonic field operators for atoms ($\s=1$) and molecules ($\s=2$),
150: $\hat{h}_{\s} = -(\hbar^2/2m_\s) \nabla^2 - \mu_\s$ are the
151: corresponding single particle Hamiltonians (focusing for
152: concreteness on the case of a homogeneous trap) with effective
153: chemical potentials $\mu_{1} = \mu$ and $\mu_2 = 2\mu - \nu$.
154: Chemical potential $\mu$ tunes the average {\em
155: total} number of atoms (whether free or bound into molecules) to
156: $N$, and detuning $\nu$ is related to the energy of a molecule at rest,
157: that can be experimentally controlled with a magnetic field. In
158: the dilute gas limit $g_1, g_2, g_{12}$ are proportional to the
159: 2-body s-wave atom-atom, atom-molecule and molecule-molecule
160: scattering lengths, respectively, and $\alpha$ characterizes
161: \emph{coherent} atom-molecule interconversion rate,
162: encoding that molecules are composed of two atoms\cite{foot1}.
163:
164: The mean-field phase diagram as a function of $\mu_{1,2}$ and
165: $\beta = 1/k_B T$ can be worked out by
166: minimizing the imaginary-time ($\tau$)
167: coherent-state action $S = \int_0^{\beta\hbar} d\tau \int d^d{\bf x}
168: \sum_{\s=1}^2 \left[\psi^*_\s\hbar \partial_\tau \psi_\s +
169: H_\mu(\psi^*_\s,\psi_\s) \right]$. Simple analysis leads to three
170: thermodynamically distinct phases (Fig.\ \ref{phasediagramA}): (i)
171: ``normal'' (N): $\Psi_{10} \equiv \langle \hat \psi_1 \rangle =
172: 0$, $\Psi_{20} \equiv \langle \hat \psi_2 \rangle = 0$, (ii)
173: ``molecular superfluid'' (MSF): $\Psi_{10} = 0, \Psi_{20} \neq 0$,
174: (iii) ``atomic superfluid'' (ASF): $\Psi_{10} \neq 0, \Psi_{20}
175: \neq 0$. Condensed atoms cause $\alpha$ to act as an effective
176: field on the molecular order parameter $\Psi_{20}$, so an
177: equilibrium phase in which atoms are condensed, but molecules are
178: not, is forbidden \cite{comment1}.
179:
180: We now examine in more detail these phases and corresponding phase
181: transitions. Phase N is stable for $\mu_{1,2} < 0$, with $\mu$
182: determined by the total atom constraint $n = n_1 + 2 n_2$, which
183: in the non-interacting limit, appropriate to a dilute weakly
184: interacting gas, is given by:
185: \begin{equation}
186: n = \frac{1}{\Lambda_T^d} \left[f_{d/2}\left(e^{\beta\mu} \right)
187: + 2^{(d+2)/2} f_{d/2}\left(e^{\beta(2\mu - \nu)} \right) \right],
188: \label{n_conserve}
189: \end{equation}
190: where $\Lambda_T = h/\sqrt{2\pi m_1 k_BT}$ is the thermal de
191: Broglie wavelength and $f_\alpha(z) = \sum_{n=1}^\infty
192: z^n/n^\alpha$ ($|z| < 1$) is the extended zeta-function.
193:
194: The N-ASF transition line $T_{c1}(n,\nu)$ occurs at $\mu = 0$ for
195: $\nu > 0$, while the N-MSF line $T_{c2}(n,\nu)$ occurs at $\mu -
196: 2\nu = 0$ for $\nu < 0$ (see Fig.\ \ref{phasediagramB}). Using
197: the appropriate asymptotics of $f_\alpha(z)$, one obtains from
198: (\ref{n_conserve}):
199: \begin{equation}
200: T_{c\sigma}(n,\nu) =
201: \begin{cases}
202: T_{c0} \left[1 + a_\sigma \left(\frac{|\nu|}{k_B T_{c0}}
203: \right)^{\frac{d-2}{2}} \right],
204: & |\nu| \ll k_B T_{c0} \\
205: T_{c\sigma}^\infty = b_\sigma c^{2/d} T_{c0},
206: & |\nu| \gg k_B T_{c0},
207: \end{cases}
208: \label{Tc}
209: \end{equation}
210: with $c = 1 + 2^{(d+2)/2}$, $a_1 = 2^{(d+4)/2}
211: |\Gamma\left(\frac{2-d}{2}\right)|/d c\zeta(d/2)$, $a_2 = 2^{-d}
212: a_1$, $b_1 = 1$, $b_2 = 2^{-(d+2)/d}$ and $T_{c0} = (h^2/2\pi m_1
213: k_B) [n/c\zeta(d/2)]^{2/d}$ the transition temperature at the
214: tricritical point $\nu = 0$.
215:
216: In the neighborhood of $T_{c1}$ the ``massive'' molecular field
217: $\hat{\psi}_2$ decouples at low energies (can be safely integrated out
218: of the partition function, leading to an effective quartic
219: coupling $g_1 \rightarrow \bar g_1 \equiv g_1 -
220: 2\alpha^2/|\mu_2|$), and the N-ASF transition is identical to that
221: of a single-component system, continuous so long as $\bar g_1 >
222: 0$. At $T=0$, the N-ASF transition takes place at vanishing atom
223: density (the N phase is simply a vacuum of atoms), and although
224: nontrivial, is exactly soluble \cite{uzunov}, corresponding to a
225: build-up of atomic superfluid (with condensate density $n_{10} =
226: |\Psi_{10}|^2 \sim |\mu|^{2\beta}$, with mean-field result $\beta
227: = 1/2$ for $d > 2$, and $\beta = d/4$ for $d < 2$) as the trap is
228: loaded. At $T \neq 0$ the N-ASF transition lies in the usual
229: $d$-dimensional XY-universality class \cite{foot2}.
230:
231: Similarly, in the neighborhood of $T_{c2}$, $\hat \psi_1$
232: decouples and the resulting N-MSF transitions are in the same
233: universality classes discussed above. The full phase boundary is
234: illustrated in Fig.\ref{phasediagramB}. In 3d it exhibits a
235: square-root singularity at the tricritical point and for $|\nu|
236: \to \infty$ asymptotes to the single-component BEC temperatures
237: $T_{c\sigma}^\infty$.
238: %with $T_{c1}^\infty/T_{c2}^\infty = 2^{5/3}$.
239:
240: \begin{figure}[tbh]
241: \centering
242: \setlength{\unitlength}{1mm}
243: \begin{picture}(70,50)(0,0)
244: \put(-22,-66){\begin{picture}(50,50)(0,0)
245: \special{psfile=phase_diagramB.eps vscale=56 hscale=56}
246: \end{picture}}
247: \end{picture}
248: \caption{Phase diagram for a bosonic atom-molecule mixture in
249: $d=3$, expressed in terms of
250: detuning $\nu$ and temperature $T$. It illustrates a finite-T
251: tricritical point at $T_{c0}$ and a quantum critical point at
252: $\nu_c(n,0)$. In the weakly-interacting limit appropriate to
253: experiments the ratio $T_{c1}^\infty/T_{c2}^\infty = 2^{5/3}$.}
254: \label{phasediagramB}
255: \end{figure}
256:
257: To study the MSF phase, we separate $\hat \psi_\sigma =
258: \Psi_{\sigma 0} + \hat{\phi}_\sigma$ into classical condensate
259: fields $\Psi_{\sigma 0}$ (with $\Psi_{10} = 0$ inside MSF) and
260: fluctuations about it. Within MSF it is sufficient to
261: expand $\hat H_\mu$ to second order in fluctuations $\hat
262: \phi_\sigma$, which leads to $\hat H_\mu = E^{(0)}(\Psi_{20}) +
263: \hat H^{(2)}$ with:
264: \begin{eqnarray}
265: E^{(0)} &=& \int d^d{\bf x}
266: \left(\Psi_{20}^* \hat h_2 \Psi_{20}
267: + \frac{g_2}{2} |\Psi_{20}|^4 \right)
268: \\
269: {\hat H}^{(2)} &=& \int d^d{\bf x} \left[\sum_{\sigma = 1}^2
270: \hat \phi_\sigma^\dagger \tilde h_\sigma \hat \phi_\sigma
271: + \frac{1}{2} \left(\lambda_\sigma \hat \phi_\sigma^\dagger
272: \hat \phi_\sigma^\dagger + {\rm h.c} \right) \right],
273: \nonumber
274: \label{Htr}
275: \end{eqnarray}
276: in which $\tilde h_\sigma = \hat h_\sigma + r_\sigma$,
277: $r_1 = g_{12} |\Psi_{20}|^2$, $r_2 = 2g_2 |\Psi_{20}|^2$, $\lambda_1 = -2\alpha
278: \Psi_{20}$, and $\lambda_2 = g_2 \Psi_{20}^2$. The linear term in
279: $\hat \phi_2, \hat \phi_2^\dagger$ vanishes automatically by the
280: self-consistent choice of $\Psi_{20}$ as the true minimum of the
281: free energy. To lowest order this gives
282: \begin{equation}
283: n_{20} \equiv |\Psi_{20}|^2 = (2\mu - \nu)/g_2,
284: \label{n20}
285: \end{equation}
286: which coincides with the minimum of $E^{(0)}$ and allows us to
287: eliminate $\mu$ in favor of $\nu$ and $n_{20}$.
288:
289: For a homogeneous system, $\hat{H}^{(2)}$ may be diagonalized by
290: Fourier transformation $\hat \phi_\sigma = V^{-1/2} \sum_{\kv}
291: e^{i\kv \cdot \xv} \hat a_{\sigma \kv}$, followed by independent
292: Bogoliubov transformations on atoms and molecules to new boson
293: operators $\gamma_{\s,\kv}, \gamma^\dagger_{\s,\kv}$:
294: \begin{eqnarray}
295: \hat{\gamma}_{\s,\kv} &=& u_{\s,\kv} \hat a_{\s,\kv} +
296: v_{\s,\kv} \hat a^\dagger_{\s,-\kv}
297: \nonumber \\
298: |u_{\s,\kv}|^2 &=& 1 + |v_{\s,\kv}|^2 = \frac{1}{2}
299: \left(\frac{\tilde \eps_{\s,\kv}}{E_{\s,\kv}} + 1 \right)
300: \nonumber \\
301: \hat{H}^{(2)} &=& \sum_{\sigma,\kv} E_{\s,\kv}
302: \left(\hat \gamma^\dagger_{\s,\kv} \hat \gamma_{\s,\kv}
303: - |v_{\s,\kv}|^2 \right),
304: \end{eqnarray}
305: in which $\tilde \eps_{\s,\kv} = \eps_{\s,\kv} - \mu_\s + r_\s$,
306: $\eps_{\s,\kv} = \hbar^2k^2/2m_\sigma$ and $E_{\s,\kv} =
307: \sqrt{\tilde \eps_{\s,\kv}^2 - |\lambda_\sigma|^2}$, and
308: $v_{\s,\kv},u_{\s,\kv}^*$ have the same phase as $\lambda_\sigma$.
309: Using (\ref{n20}) one obtains
310: \begin{eqnarray}
311: E_{1,\kv} &=& \sqrt{\left[\eps_{1,\kv} - \nu/2
312: - (g_2/2 - g_{12}) n_{20} \right]^2 - 4 \alpha^2 n_{20}}\;,
313: \nonumber \\
314: E_{2,\kv} &=& \sqrt{\eps_{2,\kv}^2 + 2g_2 n_{20}\, \eps_{2,\kv}}\;.
315: \end{eqnarray}
316: The MSF ground state is defined as vacuum of atomic and molecular
317: Bogoliubov quasi-particles $\gamma_{\s,\kv}|{\rm MSF}\rangle = 0$.
318: It is easy to show that it is given by:
319: \begin{equation}
320: |{\rm MSF}\rangle=e^{\Psi_{20} \hat a_{2,0}^\dagger}
321: \prod_{\s,\kv} e^{-\chi_{\s,\kv} \hat a_{\s,\kv}^\dagger
322: \hat a_{\s,\kv}^\dagger}|0\rangle,
323: \end{equation}
324: which is a coherent state in the $\kv = 0$ molecular mode and
325: shows BCS-like pairing correlations between time-reversed ($\kv,
326: -\kv$) atomic and molecular single particle states. The amplitude
327: $\chi_{\s,\kv} = v_{\s,\kv}/u_{\s,\kv}$ is the Fourier transform
328: of the wavefunction of $\kv \neq 0$ atom ($\sigma=1$) and molecule
329: ($\sigma=2$) pairs.
330:
331: As in a single-component SF the relation (\ref{n20}) (and more
332: fundamentally, the Goldstone theorem \cite{ZinnJustin}) ensures
333: that MSF exhibits a gapless sound mode $E_{2,\kv} \approx \hbar
334: v_0 k$, for small $k$, with $v_0 = \sqrt{g_2 n_{20}/2m_1}$,
335: corresponding to collective, long wavelength oscillations of the
336: molecular condensate. The resulting $|\kv|$ singularity in
337: $\chi_{2,\kv}$ induces a long ranged power law tail in the Fourier
338: transform $\chi_2(r) \sim 1/r^{d+1}$. In addition, we find a
339: \emph{gapped} excitation branch $E_{1,\kv}$ describing atomic-like
340: excitations, whose spectrum and eigenmodes ($\hat{\gamma}_{1,\kv},
341: \hat{\gamma}_{1,\kv}^\dagger$ do not even carry a definite atom
342: number) are qualitatively distinct from unpaired atomic
343: excitations, $\hat{a}_{1,\kv}, \hat{a}_{1,\kv}^\dagger$ of the N
344: phase. The condition of being in the MSF is that the atomic gap
345: $E^{(1)}_{\rm gap} = \sqrt{\eps_+ \eps_-}$, where
346: \begin{equation}
347: \eps_{\pm} \equiv -\nu/2-(g_2/2 - g_{12})n_{20}
348: \pm 2\alpha \sqrt{n_{20}},
349: \label{gap}
350: \end{equation}
351: remains positive. Correspondingly, the atomic pair wavefunction
352: $\chi_1(r) \sim e^{-r/\xi_1}$ decays exponentially at large
353: distance. The correlation length $\xi_1 = \hbar/\sqrt{2m \eps_-}$
354: characterizes atom-pair size, and diverges as
355: $E^{(1)}_{\rm gap}$ vanishes. The MSF-ASF transition takes place when
356: the size of atom-pairs becomes comparable to the intermolecular
357: separation, and $\xi_1$ is the diverging size of the coherent
358: exchange loops of atoms between overlapping pairs.
359:
360: In the present quadratic approximation, the free energy is given
361: by
362: \begin{eqnarray}
363: f &=& -\mu_2|\Psi_{20}|^2 + \frac{g_2}{2}|\Psi_{20}|^4
364: - \frac{1}{2V} \sum_{\s,\kv}(\tilde{\eps}_{\s,\kv} - E_{\s,\kv})
365: \nonumber\\
366: &&+\ \frac{k_B T}{V}\sum_{\s,\kv}\ln(1-e^{-E_{\s,\kv}/k_BT}).
367: \end{eqnarray}
368: The condition of fixed density $n = -(\partial f/\partial
369: \mu)_{n_{20}}$,
370: \begin{equation}
371: n = 2n_{20} + \frac{1}{V} \sum_{\s,\kv} \s
372: \left(v_{\s,\kv}^2 + \frac{u_{\s,\kv}^2 + v_{\s,\kv}^2}
373: {e^{E_{\s,\kv}/{k_BT}}-1} \right),
374: \end{equation}
375: where the first term under the summation represents the $T=0$
376: interaction-induced condensate depletion, may be used to determine
377: the condensate density $n_{20}(n,\nu,T)$. At $T=0$, in the
378: weakly-interacting limit ($a^3 n\ll 1$), in 3d we find:
379: \begin{eqnarray}
380: n_{20} &=& \frac{1}{2} n - \frac{2^{(d-2)/2}
381: \Gamma\left(\frac{d-1}{2}\right)
382: \Gamma\left(\frac{4-d}{2}\right)}{d \sqrt{\pi}
383: \Gamma\left(\frac{d}{2}\right)} (n a_2)^{d/2}
384: \nonumber \\
385: &&-\ \frac{[(n_\nu - n) \bar a_{12}]^{d/2}}
386: {2^{d+1} \Gamma\left(\frac{d}{2}\right)}
387: {\cal I}_d\left(\frac{\bar \alpha \sqrt{n}}{n_\nu - n} \right)
388: \label{n20n}
389: \end{eqnarray}
390: where $\bar a_{12} \equiv a_{2} - 2 a_{12}$,
391: $a_{2}, a_{12}$ are molecule-molecule and atom-molecule
392: scattering lengths, respectively,
393: defined by $g_{12,2} = 4\pi\hbar^2 a_{12,2}/m_2$,
394: the ``detuning density'' is $n_\nu = m_1 |\nu|/(\pi \hbar^2 \bar
395: a_{12})$, $\bar \alpha = 2^{3/2} m_1 |\alpha|/(\pi\hbar^2 \bar
396: a_{12})$, and ${\cal I}(y) = y^2 \int_0^\infty dx
397: x^{d/2}[(1+x)^2 - y^2]^{-3/2}$ is a scaling function describing
398: additional molecular-condensate depletion due to atom pairs. Equation
399: (\ref{n20n}), together with the vanishing of $\eps_-$, determines the
400: MSF-ASF phase boundary. At $T=0$, approximating $n_{20} \approx n/2$,
401: one finds:
402: \begin{equation}
403: \nu_c(n) \approx -(g_2/2-g_{12})n - 2\alpha\sqrt{2n},
404: \label{nuc}
405: \end{equation}
406: which is illustrated for $g_2 > 2g_{12}$ in Fig.\ \ref{phasediagramA}.
407: At finite $T$, in the noninteracting limit, $v_{\s,\kv} = 0$ and
408: $u_{\s,\kv} = 1$, molecular condensate density reduces to
409: \begin{equation}
410: n_{20}(\nu,T) = \frac{n}{2} \left[1 -
411: \left(\frac{T}{T_{c2}^\infty} \right)^{3/2} \right]
412: - \frac{1}{2 \Lambda_T^3} f_{3/2}\left(e^{-|\nu|/2k_B T} \right),
413: \end{equation}
414: with the standard BEC result (first term) corrected by depletion
415: due to thermal atomic excitations (last term), that is exponentially
416: small at large $|\nu|$ and low $T$.
417:
418: From the structure of the interconversion ($\alpha$) term in $\hat
419: H_\mu$ it is clear that it is a $Z_2$ (Ising) symmetry that is
420: broken at the MSF$\rightarrow$ASF transition \cite{comment2}. We
421: now show that for a homogeneous trap, both the $T=0$ and finite
422: $T$ MSF-ASF transitions lie in the Ising universality class. This
423: can be most easily seen from the coherent state action, which,
424: when expressed in terms of real and imaginary parts of the atomic
425: field $\hat \psi_1 = \hat \psi_R + i \hat \psi_I$, in MSF phase
426: reduces to:
427: \begin{eqnarray}
428: S &=& \int_0^{\beta\hbar} d\tau \int d^d{\bf x}
429: \bigg[-2i \psi_I\hbar \partial_\tau\psi_R
430: + \psi_R(\hat{h}_1 - 2\alpha\Psi_{20})\psi_R
431: \nonumber \\
432: &&\ \ \ \ \ \ \ \ \ \
433: +\ \psi_I(\hat{h}_1 + 2\alpha\Psi_{20})\psi_I\bigg]
434: + S_{\rm int}
435: \end{eqnarray}
436: where $S_{\rm int}$ are terms not essential for our argument. In
437: this form it is clear that in the presence of the molecular
438: condensate, $|\Psi_{20}| > 0$, positive $\alpha$ reduces the
439: $O(2)$ symmetry to $Z_2$, with $\psi_R$ reaching criticality {\em
440: before} $\psi_I$. Because canonically conjugate field $\psi_I$
441: remains ``massive'' at the critical point [defined by where the
442: coefficient of $\psi_R^2$ vanishes, consistent with (\ref{nuc})],
443: it can be safely integrated out and leads to a $d+1$-dimensional
444: (Lorentz-invariant) action even in the scalar order parameter
445: $\psi_R$. Therefore, as asserted above, the $T=0$ MSF-ASF
446: transition is in the $(d+1)$-dimensional Ising universality class.
447: The Ising transition is well studied, and leads to the following
448: predictions \cite{ZinnJustin}. For $d=3$, up to logarithmic
449: corrections, the mean-field theory derived above will be accurate.
450: On the other hand in 2d, MSF-ASF exponents are nontrivial but are
451: well-known. For example, standard scaling arguments predict:
452: \begin{eqnarray}
453: n_{10} \sim |\nu - \nu_c|^{2\beta_I},\ \
454: E_{gap}^{(1)} \sim |\nu - \nu_c|^{z_I \nu_I},
455: \end{eqnarray}
456: where $\beta_I \approx 0.31$, $z_I = 1$, and $\nu_I \approx 0.63$
457: are 3d Ising exponents. These, together with the
458: relevance of $T$ at this quantum critical point also imply a {\em
459: universal} shape of the MSF-ASF phase boundary $\nu_c(n,T) \sim
460: \nu_c(n,0) + a\, T^{1/\nu_I}$ in Fig.\ \ref{phasediagramB}. One
461: may hope that when long-lived molecular condensates are produced,
462: nontrivial behavior of $E_{gap}^{(1)}(\nu)$ maybe observed in
463: Ramsey fringes experiments \cite{donley}.
464:
465: Although from symmetry point of view ASF state is quite
466: conventional, it also exhibits a set of interesting features not
467: found in a single-component SF. For example, because it is a
468: discrete (Ising) symmetry that distinguishes ASF from MSF, it will
469: exhibit a gapped (Ising) mode in addition to the gapless
470: second-sound mode. As discussed above \cite{comment1} for $\alpha
471: \neq 0$, finite $\Psi_{20}$ is always induced within the ASF.
472: However, in the limit of small $\alpha$, we predict a sharp
473: crossover (indicated by a dashed $\mu_2^{\rm eff} \equiv \mu_2 -
474: \mu_1 g_{12}/\bar g_1 = 0$ curve in Figs.\ \ref{phasediagramA},
475: \ref{phasediagramB}, where the $\psi_2$ field would order for
476: $\alpha = 0$\ \cite{comment1})
477: between small molecular condensate $n_{20} \approx
478: (\alpha \mu_1/g_1 \mu_2^{\rm eff})^2$ for $\mu_2^{\rm eff} < 0$
479: and large $n_{20} \approx \mu_2^{\rm eff}/g_2$ for $\mu_2^{\rm
480: eff} > 0$, and vanishes as $n_{20} \sim (\alpha
481: n_{10})^{2/\delta}$ for $\mu_2^{\rm eff} = 0$ and weak $\alpha$,
482: with $\delta$ a universal XY-model magnetic field exponent
483: ($\delta_{d=3} \simeq 5$) \cite{ZinnJustin}.
484: %This is in direct analogy to a crossover
485: %in a weak magnetic field between high
486: %$T$, low magnetization and low $T$, high
487: %magnetization regimes in a ferromagnet.
488:
489: Another experimentally interesting feature of ASF is that a
490: seemingly standard $2\pi$ vortex in atomic condensate $\Psi_{10}$
491: will generically split into two $\pi$ vortices (see Fig.\
492: \ref{vortex_pi}) confined by a domain wall of length
493: \begin{equation}
494: r_0 = \sqrt{\hbar^2 n_{20}^{3/2}/2m\alpha n_{10}^2},
495: \end{equation}
496: that diverges as ASF$\rightarrow$MSF phase boundary is approached
497: \cite{LRunpublished}.
498: %
499: \begin{figure}[bth]
500: \centering
501: \setlength{\unitlength}{1mm}
502: \begin{picture}(70,22)(0,0)
503: \put(-35,-92){\begin{picture}(50,20)(0,0)
504: \special{psfile=vortex_pi.eps vscale=65 hscale=60}
505: \end{picture}}
506: \end{picture}
507: \caption{$2\pi$ atomic condensate vortex in ASF splits into
508: $\pi+\pi$ vortex pair connected by a ``normal'' domain wall.}
509: \label{vortex_pi}
510: \end{figure}
511: %
512: This arises because in large $\alpha$ limit a $2\pi$ vortex in atomic
513: condensate induces a $4\pi$ vortex in a molecular condensate. Such
514: double molecular vortex is unstable to two fundamental $2\pi$
515: molecular vortices, that, in 2d repel logarithmically, but are confined
516: linearly inside the ASF phase.
517:
518: Finally, standard renormalization-group analysis shows that
519: upper-critical dimensions for the tricritical point (dominated by
520: scaling of $\alpha$) are $d_{uc}^{T > 0}=6$ and $d_{uc}^{T=0} =
521: 4$, and therefore should display nontrivial critical properties
522: \cite{LRunpublished}.
523:
524: We plan to present analysis of these problems as well as
525: generalization to experimentally relevant harmonic trap in a future
526: publication.
527:
528: To conclude, we have studied a mixture of bosonic atoms and their
529: diatomic molecules and predicted an existence of a molecular
530: superfluid phase, qualitatively distinct from the normal and atomic
531: superfluid states. Ising transition MSF-ASF should be observable in
532: such systems as a function of detuning $\nu$, temperature and/or
533: density.
534:
535: We acknowledge support by NSF DMR-0321848 and Packard Foundation
536: (LR,JP), and thank Chris Greene for discussions and Steve Girvin for
537: comments on the manuscript.
538: \vspace{-0.5cm}
539: \begin{thebibliography}{99}
540: \vspace{-0.5cm}
541:
542: \bibitem{JILAbec} M. H. Anderson, {\em et al.}, {\em Science} {\bf
543: 269}, 198 (1995).
544:
545: \bibitem{MITbec} K. B. Davis {\em et al.}, \prl {\bf 75}, 3969
546: (1995).
547:
548: \bibitem{donley} E. A. Donley, {\em et al.}, {\em Nature}
549: (London) {\bf 417}, 529 (2002).
550:
551: \bibitem{jinBCS} C. A. Regal, {\em et al.}, {\em Nature} {\bf 424}, 47 (2003).
552:
553: \bibitem{foot1} The effective Hamiltonian (\ref{hamiltonian}) may
554: be derived in various limits from an underlying atomic model, in
555: which molecules are explicitly identified as bound or quasibound
556: states of atom pairs (including atomic spin dynamics, if
557: necessary); J. I. Park, L. Radzihovsky and P. B. Weichman, manuscript
558: in preparation.
559:
560: \bibitem{comment1} This contrasts with a mixture of two {\em
561: independent} species of bosons with $\alpha=0$, that in addition
562: admits a $\Psi_{10} \neq 0$, $\Psi_{20} = 0$ distinct phase.
563:
564: \bibitem{foot2} In the dilute, weakly interacting BEC limit the
565: critical behavior will be that of the ideal gas (i.e., Gaussian),
566: crossing over to XY behavior only very close to $T_{c1}$: P. B.
567: Weichman, M. Rasolt, M. J. Stephen and M. E. Fisher, Phys.\ Rev.\
568: B {\bf 33}, 4632 (1986); P. B. Weichman, Phys.\ Rev.\ B {\bf 38},
569: 8739 (1988).
570:
571: \bibitem{GradshteynRizhik} I. S. Gradshteyn and I. M. Rizhik, {\em
572: Table of Integrals, Series and Products} (Academic Press, New
573: York, 1980).
574:
575: \bibitem{uzunov} D. I. Uzunov, Phys.\ Lett.\ {\bf 87A}, 11 (1981).
576:
577: %\bibitem{sachdev} S. Sachdev, {\em Quantum Phase Transitions},
578: %Cambridge University Press, Cambridge, 1999).
579:
580: \bibitem{ZinnJustin} J. Zinn-Justin, {\em Quantum Field Theory and
581: Critical Phenomena}, (Oxford Science Puplications, Oxford, 1989).
582:
583: \bibitem{comment2} In the ``hard-spin'' $\psi_\s=e^{i\theta_\s}$
584: description, interconversion term reduces to $\alpha
585: \cos(2\theta_1-\theta_2)$, locking atomic phase to the phase of
586: the molecular condensate up to $\pi$, corresponding to the
587: unbroken $Z_2$ symmetry of the MSF. In terms of a vector order
588: parameter, MSF is characterized by a finite quadrapole moment
589: ($l=2$ angular harmonic, akin to nematic order in liquid
590: crystals), and ASF by a vector order parameter ($l=1$ harmonic).
591:
592: \bibitem{LRunpublished} L. Radzihovsky, unpublished.
593:
594: \end{thebibliography}
595:
596: \end{document}
597: