Skip to main content

The Unitary Gas and its Symmetry Properties

  • Chapter
  • First Online:
The BCS-BEC Crossover and the Unitary Fermi Gas

Part of the book series: Lecture Notes in Physics ((LNP,volume 836))

Abstract

The physics of atomic quantum gases is currently taking advantage of a powerful tool, the possibility to fully adjust the interaction strength between atoms using a magnetically controlled Feshbach resonance. For fermions with two internal states, formally two opposite spin states \(\uparrow\) and \(\downarrow\), this allows to prepare long lived strongly interacting three-dimensional gases and to study the BEC–BCS crossover. Of particular interest along the BEC–BCS crossover is the so-called unitary gas, where the atomic interaction potential between the opposite spin states has virtually an infinite scattering length and a zero range. This unitary gas is the main subject of the present chapter: it has fascinating symmetry properties, from a simple scaling invariance, to a more subtle dynamical symmetry in an isotropic harmonic trap, which is linked to a separability of the N-body problem in hyperspherical coordinates. Other analytical results, valid over the whole BEC–BCS crossover, are presented, establishing a connection between three recently measured quantities, the tail of the momentum distribution, the short range part of the pair distribution function and the mean number of closed channel molecules. The chapter is organized as follows. In Sect. 5.1, we introduce useful concepts, and we present a simple definition and basic properties of the unitary gas, related to its scale invariance. In Sect. 5.2, we describe various models that may be used to describe the BEC–BCS crossover, and in particular the unitary gas, each model having its own advantage and shedding some particular light on the unitary gas properties: scale invariance and a virial theorem hold within the zero-range model, relations between the derivative of the energy with respect to the inverse scattering length and the short range pair correlations or the tail of the momentum distribution are easily derived using the lattice model, and the same derivative is immediately related to the number of molecules in the closed channel (recently measured at Rice) using the two-channel model. In Sect. 5.3, we describe the dynamical symmetry properties of the unitary gas in a harmonic trap, and we extract their physical consequences for many-body and few-body problems.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 99.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 129.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    This toy model replaces the many-body problem with the one of a matterwave interacting with a single scatterer in a hard wall cavity of radius \(\propto 1/k_F\).

  2. 2.

    In the general case, \(\delta({\mathbf k}-{\mathbf k}^{\prime})\) has to be replaced with \(\sum_{{\mathbf K}} \delta({\mathbf k}-{\mathbf k}^{\prime}-{\mathbf K})\) where \({\mathbf K}\in (2\pi/b)\mathbb{Z}^3\) is any vector in the reciprocal lattice.

  3. 3.

    One may check, e.g. in the sector \(N_\uparrow=N_\downarrow=2,\) that the BCS variational wavefunction, which is a condensate of pairs in some pair wavefunction, does not obey the Wigner-Bethe-Peierls boundary conditions even if the pair wavefunction does, so it looses its variational character in the zero-range model.

  4. 4.

    In fixed node calculations, an everywhere negative interaction potential is used [12, 28, 29]. It is unknown if N in these simulations exceeds the minimal value required to have a bound state. Note that the imposed nodal wavefunction in the fixed node method, usually the one of the Hartree-Fock or BCS state, would be however quite different from the one of the bound state.

  5. 5.

    The inverse formula giving the full state vector in terms of \(\psi({\mathbf X})\) is \(|\Uppsi\rangle = \left(\frac{N!} {N_\uparrow!N_{\downarrow}!}\right)^{1/2} A|\!\!\uparrow\rangle^{N_\uparrow} |\!\!\!\!\downarrow\rangle^{N_\downarrow} |\psi\rangle,\) where the projector A is the usual antisymmetrizing operator \(A=(1/N!) \sum_{\sigma\in S_N} \varepsilon(\sigma) P_\sigma\).

  6. 6.

    For a continuous-space model with an interaction potential V(r), we have [76, 38] \(E_{\rm int}=\frac{C} {(4\pi)^2}\int d^3r V(r) |\phi(r)|^2\) where C is still defined by (5.63) and \(\phi(r)\) still denotes the zero-energy two-body scattering state normalised according to (5.69).

  7. 7.

    The scaling transform (5.81) defines a unitary operator \(T(\lambda)\) such that \(\psi_\lambda = T(\lambda) \psi.\) One has \(T(\lambda_1) T(\lambda_2) = T(\lambda_1 \lambda_2).\) To recover the usual additive structure as for the group of spatial translations, one sets \(\lambda=\hbox{exp}\,\,\theta\), so that \(T(\theta_1) T(\theta_2)=T(\theta_1+\theta_2)\) and \(T(\theta)=\hbox{exp}(-i\theta D)\) where D is the generator. This is why \(\ln\lambda\) appears in Eq. 5.82.

  8. 8.

    We discarded for simplicity the rather peculiar case where \(k_F |a_{\rm bg}| R_{\ast}\) is \(\le1\) but not \(\,{\ll}\,1.\)

  9. 9.

    An additional condition actually has to be imposed to have a universal gas, as we will see after Eq. 5.106.

  10. 10.

    See [68] for details. The cusp at unitarity is of course an artefact of this interpolation procedure.

  11. 11.

    This value is also compatible with the data of [77] for the one-body density matrix, whose short-range singular part is related by Fourier transformation to the large-k tail of the momentum distribution [38].

  12. 12.

    At unitarity, the local density approximation can be done analytically, yielding a relation between the contact C of the trapped gas and the one of the homogeneous gas: \(\frac{C} {N k_F^{\rm trap}}=\frac{512} {175} \frac{\zeta} {\xi^{1/4}}\) [68].

  13. 13.

    As shown in [89], \(L_+ \psi\) cannot be zero.

  14. 14.

    To obtain this inequality, one uses a virial theorem after separation of the center of mass motion [89].

  15. 15.

    For particles of equal masses one introduces the Jacobi coordinates \({\mathbf u}_i=(\frac{N-i} {N+1-i})^{1/2} \left[{\mathbf r}_i-(N-i)^{-1} \sum_{j=i+1}^{N}{\mathbf r}_j\right]\) for \(1\,{\leq}\, i\,{\leq}\, N-1.\) Then \(\Updelta_{\mathbf X} = N^{-1} \Updelta_{\mathbf C} + \sum_{i=1}^{N-1} \Updelta_{{\mathbf u}_i}\) and \(R^2=X^2-N C^{2} =\sum_{i=1}^{N-1} {\mathbf u}_i^2.\) The general case of arbitrary masses is detailed in the Appendix 3.

  16. 16.

    The reasoning below Eq. 5.143 can also be adapted by putting the center of mass in its ground state \(\psi_{\rm CM}({\mathbf C})\propto \hbox{exp}[-Nm\omega C^2/(2\hbar)]\) and by constructing purely internal raising and lowering operators of an internal SO(2,1) dynamical symmetry, that do not excite the center of mass motion contrarily to \(L_{+}\) and \(L_-\) [89].

  17. 17.

    These reformulated contact conditions are given explicitly in [55], Eq. 1.38.

  18. 18.

    Since the Fourier transform \(\tilde{A}_{ij}({\mathbf K})=\int d^{3(N-2)}Y e^{-i{\mathbf K}\cdot{\mathbf Y}} A_{ij}({\mathbf Y})\) may lead to non-absolutely converging integrals at infinity, the calculation has to be performed using the language of distributions, with a regularizing factor \(e^{-\eta Y}, \eta\to 0^{+}.\)

  19. 19.

    Strictly speaking, such a time evolution of the wavefunction in internal hyperspherical coordinates corresponds to an internal scaling solution where the center of mass wavefunction is constant, whereas the scaling solution of Sect. 5.3.1 corresponds to a hyperradial motion in the hyperspherical coordinates \((X,{\mathbf n}).\)

  20. 20.

    In article [98], the vanishing of the bulk viscosity was deduced from the so-called general coordinate and conformal invariance, the scaling solution being unknown to its author at the time of writing (although it had been obtained in [58]). The scaling solution was recently rederived using this general coordinate and conformal invariance [99]. Several other results presented in Sects. 5.3.25.3.4 were also rederived using this field theoretical formalism ([99] and refs. therein).

  21. 21.

    Strictly speaking, it is sufficient that the Laplacian on the unit sphere together with the Wigner-Bethe-Peierls boundary conditions reexpressed in terms of hyperangles is self-adjoint, as extensively used in [7]. This is less restrictive than having the full Hamiltonian self-adjoint, since it allows for example to have a N-body Efimov effect while the \(N-1\) zero-range model is perfectly well-defined and does not experience any Efimov effect.

  22. 22.

    There is a typo in Eq. 5.10 of [98]: \(\varvec{\nabla}_i(\rho v^i \partial_i s)\) should be replaced by \(\varvec{\nabla}_i(\rho v^i s)\).

  23. 23.

    If we would neglect the position-dependence of \(\eta\) and \(\zeta,\) (5.216) would reduce to the Navier-Stokes equation.

  24. 24.

    To formalize this statement, we consider a small but still macroscopic element of the equilibrium gas of volume dV around point \(\bar{\mathbf{r}}\), with \(k_F^{-1} \,{\ll}\, dV^{1/3} \,{\ll}\, R\) where \(k_F\) is the Fermi momentum and R the Thomas-Fermi radius of the gas. We can define the density operator \(\hat{\rho}_{\rm elem}\) of this element by taking the trace of the full N-body density operator over the spatial modes outside the element. Since the gauge transform in Eq. 5.109 is local in position space, \(\hat{\rho}_{\rm elem}\) experiences the same unitary gauge transform. It would be tempting to conclude from the general formula \({\it dS} = -k_B \hbox{Tr}[\hat{\rho}_{\rm elem} \ln \hat{\rho}_{\rm elem}]\) that the entropy dS of the element is not changed by the gauge transform. This is a valid conclusion however only if the gauge transform does not bring \(\hat{\rho}_{\rm elem}\) too far from local thermal equilibrium. To check this, we split the gauge transform for a single particle of position r as \(m r^{2} \dot{\lambda}/(2 \hbar \lambda) = m \dot{\lambda}/(2 \hbar \lambda) [\bar{r}^2 + 2 \bar{\mathbf{r}}\cdot (\mathbf{r}- \bar{\mathbf{r}}) + (\mathbf{r}-\bar{\mathbf{r}})^2].\) The first term is an innocuous uniform phase shift. The second term performs a uniform shift in momentum space by the announced value \(m\mathbf{v}(\bar{\mathbf{r}},t)\). Due to Galilean invariance, this has no effect on the thermodynamic quantities of the small element, such as its temperature, its pressure, its density, its entropy. With the estimate \(\dot{\lambda}/\lambda \,{\sim}\, \omega, \bar{r}\,{\sim}\, R, m\omega R \,{\sim}\, \hbar k_F\), this second term is of order \(k_F dV^{1/3}\gg 1\), not negligible. The third term is of order \(m\omega dV^{2/3}/\hbar \,{\sim}\, N^{-1/3} k_F^2 dV^{2/3}\), negligible in the thermodynamic limit.

References

  1. Newton, R.G.: Scattering theory of waves and particles, 2nd edn. Springer, Berlin (1982)

    Google Scholar 

  2. Khuri, N.N., Martin, A., Richard, J.-M., Wu, T.T.: J. Math. Phys. 50, 072105 (2009)

    Article  MathSciNet  ADS  Google Scholar 

  3. Landau, L.D., Lifschitz, E.M.: Quantum mechanics: non-relativistic theory. Elsevier Science, Oxford (2003)

    Google Scholar 

  4. Efimov, V.N., Yad, F.: 12, 1080 (1970) [Sov. J. Nucl. Phys. 12, 589 (1971)]

    Google Scholar 

  5. Efimov, V.: Nucl. Phys. A 210, 157 (1973)

    Article  ADS  Google Scholar 

  6. Petrov, D.S.: Phys. Rev. A 67, 010703 (2003)

    Article  ADS  Google Scholar 

  7. Castin, Y., Mora, C., Pricoupenko, L.: Phys. Rev. Lett. 105, 223201 (2010)

    Article  ADS  Google Scholar 

  8. Read, M., Berry, S.: Methods of modern mathematical physics 1: functional analysis. Academic Press (1980)

    Google Scholar 

  9. Ho, T-L.: Phys. Rev. Lett. 92, 090402 (2004)

    Article  ADS  Google Scholar 

  10. Uhlenbeck, G.E., Uehling, E.A.: Phys. Rev. 39, 1014 (1932)

    Article  ADS  Google Scholar 

  11. Burovski, E., Prokof’ev, N., Svistunov, B., Troyer, M.: Phys. Rev. Lett. 96, 160402 (2006)

    Article  ADS  Google Scholar 

  12. Gezerlis, A., Gandolfi, S., Schmidt, K.E., Carlson, J.: Phys. Rev. Lett. 103, 060403 (2009)

    Article  ADS  Google Scholar 

  13. Bulgac, A., Drut, J.E., Magierski, P.: Phys. Rev. A 78, 023625 (2008)

    Article  ADS  Google Scholar 

  14. Goulko, O., Wingate, M.: Phys. Rev. A 82, 053621 (2010)

    Article  ADS  Google Scholar 

  15. Busch, T., Englert, B.G., Rzazewski, K., Wilkens, M.: Found. Phys. 28, 549 (1998)

    Article  Google Scholar 

  16. Pricoupenko, L., Castin, Y.: Phys. Rev. A 69, 051601 (2004)

    Article  ADS  Google Scholar 

  17. Petrov, D.S., Salomon, C., Shlyapnikov, G.V.: Phys. Rev. Lett. 93, 090404 (2004)

    Article  ADS  Google Scholar 

  18. Bourdel, T., Cubizolles, J., Khaykovich, L., Magalhães, K.M.F., Kokkelmans, S.J.J.M.F., Shlyapnikov, G.V., Salomon, C.: Phys. Rev. Lett. 91, 020402 (2003)

    Article  ADS  Google Scholar 

  19. Jo, G.-B., Lee, Y.-R., Choi, J.-H., Christensen, C.A., Kim, T.H., Thywissen, J.H., Pritchard, D.E., Ketterle, W.: Science 325, 1521 (2009)

    Article  ADS  Google Scholar 

  20. Zhang, J., Van Kempen, E.G.M., Bourdel, T., Khaykovich, L., Cubizolles, J., Chevy, F., Teichmann, M., Tarruell, L., Kokkelmans, S.J.J.M.F., Salomon, C.: Phys. Rev. A 70, 030702(R) (2004)

    ADS  Google Scholar 

  21. Gaebler, J.P., Stewart, J.T., Bohn, J.L., Jin, D.S.: Phys. Rev. Lett. 98, 200403 (2007)

    Article  ADS  Google Scholar 

  22. Pricoupenko, L.: Phys. Rev. Lett. 96, 050401 (2006)

    Article  ADS  Google Scholar 

  23. Jona-Lasinio, M., Pricoupenko, L., Castin, Y.: Phys. Rev. A 77, 043611 (2008)

    Article  ADS  Google Scholar 

  24. Chevy, F., van Kempen, E.G.M., Bourdel, T., Zhang, J., Khaykovich, L., Teichmann, M., Tarruell, L., Kokkelmans, S.J.J.M.F., Salomon, C.: Phys. Rev. A 71, 062710 (2005)

    Article  ADS  Google Scholar 

  25. Gaudin, M.: La fonction d’onde de Bethe. Masson, Paris (1983)

    Google Scholar 

  26. Pricoupenko, L., Olshanii, M.: J. Phys. B 40, 2065 (2007)

    Article  ADS  Google Scholar 

  27. Tonini, G., Werner, F., Castin, Y.: Eur. Phys. J. D 39, 283 (2006)

    Article  ADS  Google Scholar 

  28. Chang, S.Y., Pandharipande, V.R., Carlson, J., Schmidt, K.E.: Phys. Rev. A 70, 043602 (2004)

    Article  ADS  Google Scholar 

  29. Astrakharchik, G.E., Boronat, J., Casulleras, J., Giorgini, S.: Phys. Rev. Lett. 93, 200404 (2004)

    Article  ADS  Google Scholar 

  30. Juillet, O.: New J. Phys. 9, 163 (2007)

    Article  ADS  Google Scholar 

  31. Jonsell, S., Heiselberg, H., Pethick, C.J.: Phys. Rev. Lett. 89, 250401 (2002)

    Article  ADS  Google Scholar 

  32. Werner, F., Castin, Y.: Phys. Rev. Lett. 97, 150401 (2006)

    Article  ADS  Google Scholar 

  33. Dell’Antonio, G.F., Figari, R., Teta, A.: Ann. Inst. Henri Poincaré 60, 253 (1994)

    MathSciNet  MATH  Google Scholar 

  34. Minlos, R.: On pointlike interaction between n fermions and another particle. In: Dell’Antonio G.F., Figari R., Teta A., SISSA (Trieste) (eds.) Proceedings of the Workshop on Singular Schrödinger Operators, Trieste, 29 September–1 October 1994 (1995)

    Google Scholar 

  35. Nozières, P., Schmitt-Rink, S.: J. Low Temp. Phys. 59, 195 (1985)

    Article  ADS  Google Scholar 

  36. Greiner, M., Mandel, O., Esslinger, T., Hänsch, T.W., Bloch, I.: Nature 415, 39 (2002)

    Article  ADS  Google Scholar 

  37. Castin Y.: Simple theoretical tools for low dimension Bose gases. In: Olshanii M., Perrin H., Pricoupenko L. (eds.) Lecture notes of the 2003 Les Houches Spring School, Quantum Gases in Low Dimensions, J. Phys. IV France, vol. 116, p. 89 (2004)

    Google Scholar 

  38. Werner, F., Castin, Y.: Phys. Rev. A (2010, submitted)

    Google Scholar 

  39. Randeria, M.: Bose-Einstein Condensation. In: Griffin, A., Snoke, D.W., Stringari, S. (eds.) p. 355. Cambridge University Press, Cambridge (1995)

    Google Scholar 

  40. Engelbrecht, J.R., Randeria, M., Sá de Melo, C.: Phys. Rev. B 55, 15153 (1997)

    Article  ADS  Google Scholar 

  41. Mora, C., Castin, Y.: Phys. Rev. A 67, 053615 (2003)

    Article  ADS  Google Scholar 

  42. Pricoupenko, L., Castin, Y.: J. Phys. A 40, 12863 (2007)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  43. Castin, Y.: Talk at KITP (UCSB), 2004. http://online.itp.ucsb.edu/online/gases04/castin (2004)

  44. Blaizot, J.-P., Ripka, G.: Quantum theory of finite systems. The MIT Press, Cambridge (1986)

    Google Scholar 

  45. Nascimbène, S., Navon, N., Jiang, K.J., Chevy, F., Salomon, C.: Nature 463, 1057 (2010)

    Article  ADS  Google Scholar 

  46. Castin, Y.: Basic tools for degenerate Fermi gases. In: Inguscio M., Ketterle W., Salomon C. (eds.) Lecture notes of the 2006 Varenna Enrico Fermi School on Fermi gases, SIF (2007)

    Google Scholar 

  47. Burovski, E., Prokof’ev, N., Svistunov, B., Troyer, M.: New J. Phys. 8, 153 (2006)

    Article  ADS  Google Scholar 

  48. Georges, A.: Condensed-matter physics with light and atoms: strongly correlated cold fermions in optical lattices. In: Inguscio M., Ketterle W., Salomon C. (eds.) Lecture notes of the 2006 Varenna Enrico Fermi School on Fermi gases, SIF (2007)

    Google Scholar 

  49. Tan, S.: Ann. Phys. 323, 2952 (2008)

    Article  ADS  MATH  Google Scholar 

  50. Tan, S.: Ann. Phys. 323, 2971 (2008)

    Article  ADS  MATH  Google Scholar 

  51. Lieb, E., Liniger, W.: Phys. Rev. 130, 1605 (1963)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  52. Olshanii, M., Dunjko, V.: Phys. Rev. Lett. 91, 090401 (2003)

    Article  ADS  Google Scholar 

  53. Blatt, J.M., Weisskopf, V.F.: Theoretical nuclear physics. Wiley, New York (1952)

    Google Scholar 

  54. Fregoso, B.M., Baym, G.: Phys. Rev. A 73, 043616 (2006)

    Article  ADS  Google Scholar 

  55. Werner, F.: Ph.D. thesis of the University Pierre et Marie Curie, Paris. http://tel.archives-ouvertes.fr/tel-00285587 (2008)

  56. Castin, Y.: Coherent atomic matter waves. In: Kaiser R., Westbrook C., David F. (eds.) Lecture notes of Les Houches summer school, pp. 1–136. EDP Sciences and Springer (2001)

    Google Scholar 

  57. Huang, K.: Statistical mechanics. Wiley, New York (1963)

    Google Scholar 

  58. Castin, Y.: Comptes Rendus Physique 5, 407 (2004)

    Article  ADS  Google Scholar 

  59. Mayer, I.: Simple theorems, proofs, and derivations in quantum chemistry, p. 32. Kluwer, New York (2003)

    Google Scholar 

  60. Thomas, J.E.: Phys. Rev. A 78, 013630 (2008)

    Article  ADS  Google Scholar 

  61. Thomas, J.E., Kinast, J., Turlapov, A.: Phys. Rev. Lett. 95, 120402 (2005)

    Article  ADS  Google Scholar 

  62. Tan, S.: Ann. Phys. 323, 2987 (2008)

    Article  ADS  MATH  Google Scholar 

  63. Werner, F.: Phys. Rev. A 78, 025601 (2008)

    Article  ADS  Google Scholar 

  64. Köhler, T., Góral, K., Julienne, P.S.: Rev. Mod. Phys. 78, 1311 (2006)

    Article  ADS  Google Scholar 

  65. Timmermans, E., Tommasini, P., Hussein, M., Kerman, A.: Phys. Rep. 315, 199 (1999)

    Article  ADS  Google Scholar 

  66. Holland, M., Kokkelmans, S.J.J.M.F., Chiofalo, M.L., Walser, R.: Phys. Rev. Lett. 87, 120406 (2001)

    Article  ADS  Google Scholar 

  67. Milstein, J.N., Kokkelmans, S.J.J.M.F., Holland, M.J.: Phys. Rev. A 66, 043604 (2002)

    Article  ADS  Google Scholar 

  68. Werner, F., Tarruell, L., Castin, Y.: Eur. Phys. J. B 68, 401 (2009)

    Article  ADS  Google Scholar 

  69. Szymańska, M.H., Góral, K., Köhler, T., Burnett, K.: Phys. Rev. A 72, 013610 (2005)

    Article  ADS  Google Scholar 

  70. Petrov, D.S.: Phys. Rev. Lett. 93, 143201 (2004)

    Article  ADS  Google Scholar 

  71. Gurarie, V., Radzihovsky, L.: Ann. Phys. 322, 2 (2007)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  72. Simonucci, S., Pieri, P., Strinati, G.C.: Europhys. Lett. 69, 713 (2005)

    Article  ADS  Google Scholar 

  73. Bartenstein, M., Altmeyer, A., Riedl, S., Geursen, R., Jochim, S., Chin, C., Hecker, Denschlag, J., Grimm, R., Simoni, A., Tiesinga, E., Williams, C.J., Julienne, P.S.: Phys. Rev. Lett. 94, 103201 (2005)

    Article  ADS  Google Scholar 

  74. Partridge, G.B., Strecker, K.E., Kamar, R.I., Jack, M.W., Hulet, R.G.: Phys. Rev. Lett. 95, 020404 (2005)

    Article  ADS  Google Scholar 

  75. Braaten, E., Kang, D., Platter, L.: Phys. Rev. A 78, 053606 (2008)

    Article  ADS  Google Scholar 

  76. Zhang, S., Leggett, A.J.: Phys. Rev. A 79, 023601 (2009)

    Article  ADS  Google Scholar 

  77. Lobo, C., Carusotto, I., Giorgini, S., Recati, A., Stringari, S.: Phys. Rev. Lett. 97, 100405 (2006)

    Article  ADS  Google Scholar 

  78. Kuhnle, E.D., Hu, H., Liu, X.-J., Dyke, P., Mark, M., Drummond, P.D., Hannaford, P., Vale, C.J.: Phys. Rev. Lett. 105, 070402 (2010)

    Article  ADS  Google Scholar 

  79. Stewart, J.T., Gaebler, J.P., Drake, T.E., Jin, D.S.: Phys. Rev. Lett. 104, 235301 (2010)

    Article  ADS  Google Scholar 

  80. Contribution of E. Braaten in Chapter 6 of this volume, also available at arXiv:1008.2922

    Google Scholar 

  81. Ho, T.-L., Zhou, Q.: Nature Phys. 6, 131 (2010)

    Article  ADS  Google Scholar 

  82. Navon, N., Nascimbène, S., Chevy, F., Salomon, C.: Science 328, 729 (2010)

    Article  ADS  Google Scholar 

  83. Leggett, A.J.: Lecture 2 at the International Quantum Gases Workshop, Institut Henri Poincaré, Paris. http://www.phys.ens.fr/˜castin/files_pdf.html (2007)

  84. Lewis, H.R.: J. Math. Phys. 9, 1976 (1968)

    Article  ADS  MATH  Google Scholar 

  85. Ermakov, V.P.: Univ. Izv. Kiev. 20, 1 (1880)

    Google Scholar 

  86. Kagan, Yu., Surkov, E.L., Shlyapnikov, G.V.: Phys. Rev. A 54, R1753 (1996)

    Article  ADS  Google Scholar 

  87. Minguzzi, A., Gangardt, D.M.: Phys. Rev. Lett. 94, 240404 (2005)

    Article  ADS  Google Scholar 

  88. Lobo, C., Gensemer, S.D.: Phys. Rev. A 78, 023618 (2008)

    Article  ADS  Google Scholar 

  89. Werner, F., Castin, Y.: Phys. Rev. A 74, 053604 (2006)

    Article  ADS  Google Scholar 

  90. Pitaevskii, L.P., Rosch, A.: Phys. Rev. A 55, R853 (1997)

    Article  ADS  Google Scholar 

  91. Tan, S.: arXiv:cond-mat/0412764

    Google Scholar 

  92. Faddeev, L.D., Merkuriev, S.P.: Quantum scattering theory for several particle systems, Section 1.3.1. In: Flato M. (ed.) Kluwers Academic Publishers, Dordrecht (1993)

    Google Scholar 

  93. Skorniakov, G.V., Ter-Martirosian, K.A.: Sov. Phys. JETP 4, 648 (1957)

    MathSciNet  Google Scholar 

  94. Danilov, G.S.: Sov. Phys. JETP 13, 349 (1961)

    MathSciNet  MATH  Google Scholar 

  95. Birse, M.C.: J. Phys. A 39, L49 (2006)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  96. Blume, D., von Stecher, J., Greene, C.H.: Phys. Rev. Lett. 99, 233201 (2007)

    Article  ADS  Google Scholar 

  97. Liu, X.-J., Hu, H., Drummond, P.D.: Phys. Rev. Lett. 102, 160401 (2009)

    Article  ADS  Google Scholar 

  98. Son, D.T.: Phys. Rev. Lett. 98, 020604 (2007)

    Article  ADS  Google Scholar 

  99. Contribution by Y. Nishida and D. T. Son in Chapter 7 of this volume, also available at arXiv:1004.3597

    Google Scholar 

  100. Schäfer, T.: arXiv:1008:3876

    Google Scholar 

  101. Landau, L.D., Lifshitz, E.M.: Fluid mechanics. Pergamon, New York (1987)

    MATH  Google Scholar 

  102. Caldirola, P., Cirelli, R., Prosperi, G.: Introduzione alla fisica teorica, p. 534. Utet, Torino (1982)

    Google Scholar 

  103. Nishida, Y., Son, D.T., Tan, S.: Phys. Rev. Lett. 100, 090405 (2008)

    Article  ADS  Google Scholar 

  104. Pricoupenko, L.: Phys. Rev. A 73, 012701 (2006)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We thank O. Goulko, O. Juillet, T. Schäfer, D.T. Son, B. Svistunov and S. Tan for helpful discussions while writing this manuscript, and the authors of [21] for their data. F. W. is supported by NSF grant PHY-0653183, Y. C. is member of IFRAF and acknowledges support from the ERC project FERLODIM N.228177.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yvan Castin .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Effective Range in a Lattice Model

To calculate the effective range \(r_e\) [defined by Eq. 5.5] for the lattice model of Sect. 5.2.1, it is convenient to perform in the expression (5.46) of the scattering amplitude an analytic continuation to purely imaginary incoming wavevectors \(k_0\), setting \(k_0=i q_0\) with \(q_0\) real and positive. Eliminating \(1/g_0\) thanks to Eq. 5.47 we obtain the useful expression:

$$ -\frac{1}{f_{k_0}} = \frac{1}{a} +4\pi \int_{{\fancyscript{D}}} \frac{d^3k}{(2\pi)^3} \left[\frac{1}{q_0^2+2m\varepsilon_{\mathbf{k}}/\hbar^2}-\frac{1}{2m\varepsilon_{\mathbf{k}}/\hbar^2}\right]. $$
(5.180)

We first treat the case of the parabolic dispersion relation Eq. 5.48. A direct expansion of Eq. 5.180 in powers of \(q_0\) leads to an infrared divergence. The trick is to use the fact that the integral over \(\fancyscript{D}\) in Eq. 5.180 can be written as the integral of the same integrand over the whole space minus the integral over the supplementary space \(\mathbb{R}^3\setminus \fancyscript{D}\). The integral over the whole space may be performed exactly using

$$ \int_{{\mathbb{R}}^3} \frac{d^3k}{(2\pi)^3} \left[\frac{1}{q_0^2+k^2}-\frac{1}{k^2}\right] = -\frac{q_0}{4\pi}. $$
(5.181)

This leads to the transparent expression, where the term corresponding to ik in Eq. 5.3, and which is non-analytic in the energy E, is now singled out:

$$ -\frac{1}{f_{k_0}^{\rm parab}} = \frac{1}{a} - q_0 -4\pi \int_{{\mathbb{R}}^3\setminus{\fancyscript{D}}} \frac{d^3k}{(2\pi)^3} \left[\frac{1}{q_0^2+k^2}-\frac{1}{k^2}\right]. $$
(5.182)

This is now expandable in powers of \(q_0^{2},\) leading to the effective range for the parabolic dispersion relation:

$$ r_e^{\rm parab} = \frac{1}{\pi^2} \int_{{\mathbb{R}}^3\setminus{\fancyscript{D}}} \frac{d^3k}{k^4}. $$
(5.183)

We now turn back to the general case. The trick is to consider the difference between the inverse scattering amplitudes of the general case and the parabolic case with a common value of the scattering length:

$$ \frac{1}{f_{k_0}^{\rm parab}} - \frac{1}{f_{k_0}} = 4\pi \int_{{\fancyscript{D}}} \frac{d^3k}{(2\pi)^3} \left[\frac{1}{q_0^2+2m\varepsilon_{\mathbf{k}}/\hbar^2}-\frac{1}{q_0^2+k^2} -\frac{1}{2m\varepsilon_{\mathbf{k}}/\hbar^2} + \frac{1}{k^2}\right]. $$
(5.184)

This is directly expandable to second order in \(q_0,\) leading to:

$$ r_e - r_e^{\rm parab} = 8\pi \int_{{\fancyscript{D}}} \frac{d^3k}{(2\pi)^3} \left[ \frac{1}{k^4} - \left(\frac{\hbar^2}{2m\varepsilon_{\mathbf{k}}}\right)^{2} \right]. $$
(5.185)

The numerical evaluation of this integral for the Hubbard dispersion relation Eq. 5.59 leads to the Hubbard model effective range Eq. 5.60.

Finally, we specialize the general formula to the parabolic plus quartic form Eq. 5.61. Setting \(\mathbf{k}=(\pi/b)\mathbf{q}\) and using Eq. 5.183, we obtain

$$ \frac{\pi^3 r_e^{\rm mix}}{b} = \int_{{\mathbb{R}}^3\setminus[-1,1]^3} \frac{d^3q}{q^4} + \int_{[-1,1]^3} \frac{d^3q}{q^4} \left[1-\frac{1}{(1-Cq^2)^2}\right]. $$
(5.186)

The trick is to split the cube \([-1,1]^3\) as the union of \(B(0,1),\) the sphere of center 0 and unit radius, and of the set \(X=[-1,1]^3\setminus B(0,1).\) One has also \(\left(\mathbb{R}^3\setminus[-1,1]^3\right) \cup X=\mathbb{R}^3\setminus B(0,1)\) so that

$$ \frac{\pi^3 r_e^{\rm mix}}{b} = \int_{{\mathbb{R}}^3\setminus B(0,1)} \frac{d^3q}{q^4} +\int_{B(0,1)} \frac{d^3q}{q^4} \left[1-\frac{1}{(1-Cq^2)^2}\right] - \int_X \frac{d^3q}{q^4} \frac{1}{(1-Cq^2)^2}. $$
(5.187)

One then moves to spherical coordinates of axis z. The first two terms in the right hand side may be calculated exactly. In particular, one introduces a primitive of \(q^{-2}(1-C q^2)^{-2},\) given by \(C^{1/2} \Upphi(C^{1/2}q)\) with

$$ \Upphi(x) = \frac{x}{2(1-x^2)} + \frac{3}{2}\hbox{arctanh}\,x -\frac{1}{x}. $$
(5.188)

In the last term of Eq. 5.187 one integrates over the modulus q of q for a fixed direction \((\theta,\phi)\) where \(\theta\) is the polar angle and \(\phi\) the azimuthal angle. One then finds that q ranges from 1 to some maximal value \(Q(\theta,\phi),\) and the integral over q provides the difference \(\Upphi(C^{1/2}Q)-\Upphi(C^{1/2}).\) Remarkably, the term \(-\Upphi(C^{1/2})\) cancels the contribution of the first two integrals in the right hand side of Eq. 5.187, so that

$$ \frac{r_e^{\rm mix}}{b} = -\frac{C^{1/2}}{\pi^3} \int_0^{2\pi} d\phi \int_{-1}^{1} du\,\Upphi[C^{1/2}Q(\theta,\phi)] $$
(5.189)

where as usual we have set \(u=\cos\theta\). Using the symmetry under parity along each Cartesian axis, which adds a factor 8, and restricting to the face \(q_x=1\) of the cube, which adds a factor 3, the expression of \(Q(\theta,\phi)\) is readily obtained, leading to

$$ \frac{r_e^{\rm mix}}{b} = -\frac{24 C^{1/2}}{\pi^3} \int_0^{\pi/4} d\phi \int_0^{\displaystyle\frac{\hbox{cos}\phi}{\sqrt{1+\hbox{cos}^2\phi}}} du \,\Upphi\left(\frac{C^{1/2}}{\hbox{cos}\phi \sqrt{1-u^2}}\right). $$
(5.190)

In the limit \(C\to 0\), \(r_e^{\rm mix}\to r_e^{\rm parab}\), and Eq. 5.190 may be calculated analytically with \(\Upphi(x)\,{\sim}\, -1/x\) and with an exchange of the order of integration: This leads to Eq. 5.58. For a general value of \(C\in [0,1/3]\) we have calculated Eq. 5.190 numerically, and we have identified the magic value of C leading to a zero effective range, see Eq. 5.62. With the same technique, we can calculate the value of K appearing in Eq. 5.49 from the expression

$$ K = \frac{12}{\pi C^{1/2}} \int_0^{\pi/4} d\phi \int_0^{\displaystyle\frac{\hbox{cos}\phi}{\sqrt{1+\hbox{cos}^2\phi}}} {\it du}\,\,\hbox{arctanh}\,\frac{C^{1/2}}{\hbox{cos}\phi \sqrt{1-u^2}}. $$
(5.191)

Appendix 2: What is the Domain of a Hamiltonian?

Let us consider a Hamiltonian H represented by a differential operator also called H. A naive and practical definition of the domain D(H) of H is that it is the set of wavefunctions over which the action of the Hamiltonian is indeed represented by the considered differential operator. In other words, if a wavefunction \(\psi_{\rm bad}\) does not belong to D(H), one should not calculate the action of H on \(\psi_{\rm bad}\) directly using the differential operator H. If H if self-adjoint, one should rather expand \(\psi_{\rm bad}\) on the Hilbert basis of eigenstates of H and calculate the action of H in this basis.

For example, for a single particle in one dimension in a box with infinite walls in \(x=0\) and \(x=1\), so that \(0\,{\leq}\, x \,{\leq}\, 1\), one has the Hamiltonian

$$ H = -\frac{1}{2} \frac{d^2}{dx^2}, $$
(5.192)

with the boundary conditions on the wavefunction

$$ \psi(0)=\psi(1)=0 $$
(5.193)

representing the effect of the box. To be in the domain, a wavefunction \(\psi(x)\) should be twice differentiable for \(0<x<1\) and should obey the boundary conditions (5.193). An example of a wavefunction which is not in the domain is the constant wavefunction \(\psi(x)=1.\) An example of wavefunction in the domain is

$$ \psi(x) = 30^{1/2} x (1-x). $$
(5.194)

If one is not careful, one may obtain wrong results. Let us calculate the mean energy and the second moment of the energy for \(\psi\) given by (5.194). By repeated action of H onto \(\psi\), and calculation of elementary integrals, one obtains

$$ \langle H\rangle_\psi = 5 $$
(5.195)
$$ \langle H^2\rangle_\psi = 0 ?! $$
(5.196)

Eq. 5.195 is correct, but Eq. 5.196 is wrong (it would lead to a negative variance of the energy) because \(H\psi\) is not in D(H) and the subsequent illicit action of H as the differential operator (5.192) gives zero.

How to calculate the right value of \(\langle H^2\rangle_\psi\)? One introduces the orthonormal Hilbert basis of eigenstates of H,

$$ \psi_n(x) = 2^{1/2} \hbox{sin} [\pi(n+1)x], \quad n\in {\mathbb{N}}, $$
(5.197)

with the eigenenergy \(\varepsilon_n=\frac{\pi^2}{2} (n+1)^2\). Then \(\psi\) of Eq. 5.194 may be expanded as \(\sum_n c_n \psi_n(x)\), and the \(k{\rm th}\) moment of the energy may be defined as

$$ \langle H^k\rangle_\psi = \sum_{n\in {\mathbb{N}}} (\varepsilon_n)^k |c_n|^2. $$
(5.198)

Since \(c_n=4\sqrt{15}[1+(-1)^n]/[\pi(n+1)]^3,\) one recovers \(\langle H\rangle_\psi =5\) and one obtains the correct value \(\langle H^2\rangle_\psi=30,\) that leads to a positive energy variance as it should be. Also \(\langle H^k\rangle_\psi=+\infty\) for \(k\,{\geq}\, 3.\)

The trick of expanding \(\psi\) in the eigenbasis of H is thus quite powerful, it allows to define the action of H on any wavefunction \(\psi\) in the Hilbert space (not belonging to the domain). It may be applied of course only if H is self-adjoint, as it is the case in our simple example.

Appendix 3: Separability and Jacobi Coordinates for Arbitrary Masses

We here consider \(N\,{\geq}\, 2\) harmonically trapped particles interacting in the unitary limit, with possibly different masses \(m_i\) but with the same isotropic angular oscillation frequency \(\omega.\) The Hamiltonian reads

$$ H = \sum_{i=1}^{N} \left[ -\frac{\hbar^2}{2m_i} \Updelta_{{\mathbf{r}}_i} + \frac{1}{2} m_i \omega^2 r_i^2 \right] $$
(5.199)

and the unitary interaction is described by the Wigner-Bethe-Peierls contact conditions on the N-body wavefunction: For all pairs of particles (i, j), in the limit \(r_{ij}=|\mathbf{r}_i-\mathbf{r}_j|\to 0\) with a fixed value of the centroid of the particles i and j, \(\mathbf{R}_{ij}\equiv (m_i\mathbf{r}_i+m_j \mathbf{r}_j)/(m_i+m_j),\) that differs from the positions \(\mathbf{r}_k\) of the other particles, \(k\neq i,j,\) there exists a function \(A_{ij}\) such that

$$ \psi({\mathbf{r}}_1,\ldots,{\mathbf{r}}_N) = \frac{A_{ij}({\mathbf{R}}_{ij};({\mathbf{r}}_k)_{k\neq i,j})}{r_{ij}} + O(r_{ij}). $$
(5.200)

As is well known and as we will explain below, the internal Hamiltonian \(H_{\rm internal}= H - H_{\rm CM}\), where \(H_{\rm CM}= -\frac{\hbar^2}{2M}\Updelta_\mathbf{C} + \frac{1}{2} M \omega^2 C^2,\) takes the form

$$ H_{\rm internal} = \sum_{i=1}^{N-1} \left[ -\frac{\hbar^2}{2\bar{m}} \Updelta_{{\mathbf{u}}_i} + \frac{1}{2} \bar{m} \omega^2 u_i^2 \right] $$
(5.201)

in suitably defined Jacobi coordinates [see Eqs. 5.205, 5.214]. Here \(\mathbf{C}=\sum_{i=1}^{N} m_i\mathbf{r}_i/M\) is the center of mass position, \(M= \sum_{i=1}^{N} m_i\) is the total mass, and \(\bar{m}\) is some arbitray mass reference, for example the mean mass \(M/N\). Then it is straightforward to express Eq. 5.201 in hyperspherical coordinates, the vector \((\mathbf{u}_1,\ldots,\mathbf{u}_{N-1})\) with \(3N-3\) coordinates being expressed in terms of its modulus R and a set of \(3N-4\) hyperangles \(\Upomega,\) so that

$$ H_{\rm internal} = -\frac{\hbar^2}{2\bar{m}} \left[\partial_R^{2} + \frac{3N-4}{R} \partial_R +\frac{1}{R^2} \Updelta_{\Upomega} \right] +\frac{1}{2} \bar{m} \omega^{2} R^{2} $$
(5.202)

where \(\Updelta_{\Upomega}\) is the Laplacian over the unit sphere of dimension \(3N-4.\) As we shall see, the expression for the hyperradius is simply

$$ R^{2} \equiv \sum_{i=1}^{N-1} u_i^{2} = \frac{1}{\bar{m}} \sum_{i=1}^{N} m_{i} ({\mathbf{r}}_i - {\mathbf{C}})^{2}. $$
(5.203)

This form of the Hamiltonian is then useful to show the separability of Schrödinger’s equation for the unitary gas in hyperspherical coordinates [55, 89] for \(N\,{\geq}\, 3\) and arbitrary masses. The separability Eq. 5.161 that was described for simplicity in the case of equal mass particles in Sect. 5.3.3 indeed still holds in the case of different masses, if the Wigner-Bethe-Peierls model defines a self-adjoint Hamiltonian. Footnote 21 We recall here the various arguments. First, for zero energy free space eigenstates, the form Eq. 5.159 is expected from scale invariance, if the Hamiltonian is self-adjoint [89]. Second, the form Eq. 5.161 for the general case, including non-zero energy and an isotropic harmonic trap, is expected because (i) the Hamiltonian (5.199), after separation of the center of mass, has the separable form (5.202) in hyperspherical coordinates, and (ii) Eq. 5.161 obeys the Wigner-Bethe-Peierls contact conditions if Eq. 5.159 does. This point (ii) results from the fact that the Wigner-Bethe-Peierls conditions are imposed, for each pair of particles (i,j), for \(r_{ij}\to 0\) with a fixed value of \(\mathbf{R}_{ij}\) that differs from the positions \(\mathbf{r}_k\) of the other particles, \(k\neq i,j.\) Using \(\mathbf{r}_i=\mathbf{R}_{ij}+[m_j/(m_i+m_j)] \mathbf{r}_{ij}\) and \(\mathbf{r}_j=\mathbf{R}_{ij}-[m_i/(m_i+m_j)] \mathbf{r}_{ij}\), with \(\mathbf{r}_{ij}\equiv \mathbf{r}_i-\mathbf{r}_{j}\), we indeed find that

$$ \bar{m} R^2 = \frac{m_i m_j}{m_i+m_j} r_{ij}^2 + (m_i+m_j) ({\mathbf{R}}_{ij}-{\mathbf{C}})^2 + \sum_{k\neq i,j} m_k ({\mathbf{r}}_k-{\mathbf{C}})^2. $$
(5.204)

For \(N\,{\geq}\, 3\), we see that \(\lim_{r_{ij}\to 0} R^2 > 0,\) so that R varies only to second order in \(r_{ij}\) in that limit. Provided that the function F(R) in Eq. 5.161 has no singularity at non-zero R, the Wigner-Bethe-Peierls contact conditions are preserved [similarly to the argument Eq. 5.115]. Third, bosonic or fermionic exchange symmetries imposed on the N-body wavefunction cannot break the separability in hyperspherical coordinates: Exchanging the positions of particles of same mass does not change the value of the hyperradius R, it only affects the hyperangles and thus the eigenvalues \([(3N-5)/2]^2-s^{2}\) of the Laplacian on the unit sphere.

To derive the form Eq. 5.201 of the internal Hamiltonian, we introduce the usual Jacobi coordinates given for example in [102]:

$$ {\mathbf{y}}_i \equiv {\mathbf{r}}_i -\frac{\sum_{j=i+1}^{N} m_j {\mathbf{r}}_j}{\sum_{j=i+1}^{N} m_j} \quad \hbox{for } 1\,{\leq}\, i \,{\leq}\, N-1. $$
(5.205)

We note that \(\mathbf{y}_i\) simply gives the relative coordinates of particle i with respect to the center of mass of the particles from \(i+1\) to N. To simplify notations, we also set \(\mathbf{y}_N \equiv \mathbf{C}\). Here we derive Eq. 5.201 in a pedestrian way. Note that a more elegant derivation by recursion is given in page 63 of [55]. In compact form, the Jacobi change of variables corresponds to setting \(\mathbf{y}_i = \sum_{j=1}^{N} M_{ij} \mathbf{r}_j\) for \(1\,{\leq}\, i \,{\leq}\, N\), where the non-symmetric matrix M is such that:

  • In the case \(1\,{\leq}\, i<N\), one has: \(M_{ij}=0\) for \(1\,{\leq}\, j < i\), \(M_{ij}=1\) for \(j=i\), and \(M_{ij}=-m_j/\left(\sum_{k=i+1}^{N} m_k\right)\) for \(i< j \,{\leq}\, N\).

  • \(M_{Nj} = m_j/\left(\sum_{k=1}^{N} m_k\right)\) for \(1 \,{\leq}\, j \,{\leq}\, N\).

From the formula giving the derivative of a composite function, the kinetic energy operator writes

$$ H_{\rm kin} \equiv \sum_{i=1}^{N} -\frac{\hbar^2}{2m_i} \Updelta_{{\mathbf{r}}_i} = -\frac{\hbar^2}{2} \sum_{j=1}^{N} \sum_{k=1}^{N} S_{jk} {\hbox{grad}}_{{\mathbf{y}}_j} \cdot {\hbox{grad}}_{{\mathbf{y}}_{k}}, $$
(5.206)

where the symmetric matrix S is defined as \(S_{jk} = \sum_{i=1}^{N} M_{ji} M_{ki}/m_i.\) The explicit calculation of the matrix elements \(S_{jk}\) is quite simple. Taking advantage of the fact that S is symmetric, one has to distinguish three cases, (i) \(1\,{\leq}\, j,k\,{\leq}\, N-1\), with \(j=k\) and \(j<k\) as subcases, (ii) \(j=k=N,\) and (iii) \(j<N, k=N.\) One then finds that \(S\) is purely diagonal, with \(S_{ii}=1/\mu_i\) for \(1\,{\leq}\, i\,{\leq}\, N-1\) and \(S_{NN}=1/M\). Here \(\mu_i\) is the reduced mass for the particle i and for a fictitious particle of mass equal to the sum of the masses of the particles from \(i+1\) to N:

$$ \frac{1}{\mu_i} = \frac{1}{m_i} + \frac{1}{\sum_{j=i+1}^{N} m_j} \quad \hbox{for} \ \ 1\,{\leq}\, i\,{\leq}\, N-1. $$
(5.207)

This results in the following form

$$ H_{\rm kin} = -\frac{\hbar^2}{2M} \Updelta_{{\mathbf{C}}} - \sum_{i=1}^{N-1} \frac{\hbar^2}{2\mu_i} \Updelta_{{\mathbf{y}}_i}. $$
(5.208)

The next step is to consider the trapping potential energy term. Inspired by Eq. 5.208 one may consider the guess

$$ H_{\rm trap} \equiv \sum_{i=1}^{N} \frac{1}{2} m_i \omega^2 r_i^2 {\mathop = \limits^{?}} \frac{1}{2} M\omega^2 C^2 + \sum_{i=1}^{N-1} \frac{1}{2} \mu_i \omega^2 y_i^2. $$
(5.209)

Replacing each \(\mathbf{y}_i\) by their expression in the guess gives

$$ M C^{2} + \sum_{i=1}^{N-1} \mu_i y_i^2 = \sum_{j=1}^{N} \sum_{k=1}^{N} Q_{jk} {\mathbf{r}}_j \cdot {\mathbf{r}}_k $$
(5.210)

where Q is uniquely defined once it is imposed to be a symmetric matrix. Setting \(M_i=\sum_{j=i+1}^{N} m_j\) for \(0\,{\leq}\, i \,{\leq}\, N-1,\) and \(M_N=0,\) we find for the off-diagonal matrix elements

$$ Q_{jk} = -\frac{\mu_{{\rm min}(j,k)}m_{{\rm max}(j,k)}}{M_{{\rm min}(j,k)}} + \frac{m_j m_k}{M} +m_j m_k \sum_{i=1}^{{\rm min}(j,k)-1} \frac{\mu_i}{M_i^2} $$
(5.211)

where \(1\,{\leq}\, j,k\,{\leq}\, N, \hbox{min}(j,k)\) and \(\hbox{max}(j,k)\) respectively stand for the smallest and for the largest of the two indices j and k. The key relation is then that

$$ \frac{\mu_i}{M_i^2} = \frac{1}{M_i} - \frac{1}{m_i+M_i} =\frac{1}{M_i} - \frac{1}{M_{i-1}} $$
(5.212)

since \(M_i + m_i = M_{i-1}\) for \(1\,{\leq}\, i\,{\leq}\, N\). This allows to calculate the sum over \(i\) of \(\mu_i/M_i^2\), as all except the border terms compensate by pairs. E.g. for \(j < k\):

$$ \sum_{i=1}^{j-1} \frac{\mu_i}{M_i^2} = \frac{1}{M_{j-1}}- \frac{1}{M} $$
(5.213)

since \(M_0=M\). One then finds that the off-diagonal elements of the matrix Q vanish. The diagonal elements of Q may be calculated using the same tricks (5.212, 5.213), one finds \(Q_{ii}= m_i\) for \(1\,{\leq}\, i\,{\leq}\, N.\) As a consequence, the guess was correct and the question mark can be removed from Eq. 5.209.

The last step to obtain Eq. 5.201 is to appropriately rescale the usual Jacobi coordinates, setting

$$ {\mathbf{u}}_i \equiv (\mu_i/\bar{m})^{1/2} {\mathbf{y}}_i $$
(5.214)

where \(\bar{m}\) is an arbitrarily chosen mass. A useful identity is the expression for the square of the hyperradius, Eq. 5.203. Starting from the definition [first identity in Eq. 5.203] we see that \(R^2=\sum_{i=1}^{N-1} \frac{\mu_i}{\bar{m}} y_i ^2.\) Then the second identity in Eq. 5.203 results from the fact that the guess in Eq. 5.209 is correct.

Appendix 4: Hydrodynamic Equations

The hydrodynamic equations for a normal compressible viscous fluid are (see [98], Footnote 22 or Sect. 15 and Sect. 49 in [101]):

  • the continuity equation

    $$ \frac{\partial \rho}{\partial t}+\varvec{\nabla}\cdot(\rho{\mathbf{v}}) =0, $$
    (5.215)
  • the equation of motion

    $$ \begin{aligned} m \rho \left(\frac{\partial {\it v}_{i} }{\partial t}+{\mathbf{v}}\cdot{\varvec\nabla} {\it v}_{i}\right)&=-\frac{\partial p}{\partial x_i}-\rho\frac{\partial U}{\partial x_i}+\sum_k\frac{\partial}{\partial x_k}\left[\eta\left(\frac{\partial{\it v}_{i}}{\partial x_k}+\frac{\partial{\it v}_{k}}{\partial x_i}-\frac{2}{3}\delta_{ik}\varvec{\nabla}\cdot{\mathbf{v}}\right)\right] \\ &\enskip\quad+\frac{\partial}{\partial x_i}\left(\zeta\,\varvec{\nabla}\cdot{\mathbf{v}}\right) \end{aligned} $$
    (5.216)

    where m is the atomic mass, \(\eta\) is the shear viscosity, \(\zeta\) is the bulk viscosity, and the pressure \(p(\mathbf{r},t)\) [as well as the temperature \(T(\mathbf{r},t)\) appearing in the next equation] is as always expressible in terms of \(\rho(\mathbf{r},t)\) and \(s(\mathbf{r},t)\) via the equation of state, Footnote 23

  • the entropy-production equation

    $$ \rho T \left(\frac{\partial s}{\partial t}+{\mathbf{v}}\cdot\varvec{\nabla}s\right) = \varvec{\nabla}\cdot(\kappa\varvec{\nabla}T)+\frac{\eta}{2}\sum_{i,k}\left( \frac{\partial {\it v}_{i}}{\partial x_k}+\frac{\partial{\it v}_{k}}{\partial x_i}-\frac{2}{3}\delta_{ik}\varvec{\nabla}\cdot{\mathbf{v}} \right)^{2} +\zeta\|\varvec{\nabla}\cdot{\mathbf{v}}\|^2 $$
    (5.217)

    where \(\kappa\) is the thermal conductivity.

Appendix 5: Alternative Derivation of the Vanishing Bulk Viscosity

Consider the particular case of a unitary gas initially prepared at thermal equilibrium in an isotropic harmonic trap at a temperature T above the critical temperature. When the harmonic trap becomes time dependent, \(U(\mathbf{r},t)=\frac{1}{2} m \omega^2(t) r^2\), each many-body eigenstate of the statistical mixture evolves under the combination Eq. 5.109 of a time dependent gauge transform and a time dependent scaling transform of scaling factor \(\lambda(t)\). The effect of the gauge transform is to shift the momentum operator \(\mathbf{p}_i\) of each particle i by the spatially slowly varying operator \(m\mathbf{r}_i \dot{\lambda}/\lambda\). In the hydrodynamic framework, this is fully included by the velocity field Eq. 5.178. Footnote 24 Using the macroscopic consequences of a spatial scaling Eqs. 5.195.22, one a priori obtains a time dependent solution of the hydrodynamic equations:

$$ T({\mathbf{r}},t) = T(t=0)/\lambda^2(t) $$
(5.218)
$$ \rho({\mathbf{r}},t) = \rho({\mathbf{r}}/\lambda,0)/\lambda^3(t) $$
(5.219)
$$ s({\mathbf{r}},t) = s({\mathbf{r}}/\lambda,0) $$
(5.220)
$$ p({\mathbf{r}},t) = p({\mathbf{r}}/\lambda,0)/\lambda^5(t) $$
(5.221)
$$ {\it v}_{i}({\mathbf{r}},t) = x_i \dot{\lambda}(t)/\lambda(t). $$
(5.222)

One then may a posteriori check that Eq. 5.215 is inconditionally satisfied, and that Eq. 5.217 is satisfied if \(\zeta\equiv 0.\) Setting \(\zeta\equiv 0\) in Eq. 5.216, and using the hydrostatic condition \({\nabla}p=-\rho{\nabla} U\) at time \(t=0,\) one finds that Eq. 5.216 holds provided that \(\lambda(t)\) solves Eq. 5.111 as it should be.

Appendix 6: \({\boldsymbol{n}}\)-Body Resonances

Usually in quantum mechanics one takes the boundary condition that the wavefunction is bounded when two particles approach each other; in contrast, the Wigner-Bethe-Peierls boundary condition (5.75) expresses the existence of a two-body resonance. If the interaction potential is fine-tuned not only to be close to a two-body resonance (i.e. to have \(|a|\gg b\)) but also to be close to a n-body resonance (meaning that a real or virtual n-body bound state consisting of \(n_\uparrow\) particles of spin \(\uparrow\) and \(n_\downarrow\) particles of spin \(\downarrow\) is close to threshold), then one similarly expects that, in the zero-range limit, the interaction potential can be replaced by the Wigner-Bethe-Peierls boundary condition, together with an additional boundary condition in the limit where any subset of \(n_\uparrow\) particles of spin \(\uparrow\) and \(n_\downarrow\) particles of spin \(\downarrow\) particles approach each other. Using the notations of Sect. 5.3.6, this additional boundary condition reads [55, 70, 89, 103]:

$$ \psi({\mathbf{r}}_1,\ldots,{\mathbf{r}}_N) = \left(R_J^{-s}-\frac{\varepsilon}{l^{2s}} R_J^s\right)\,R_J^{-\frac{3 n-5}{2}}\,\phi(\Upomega_J) \,A_J({\mathbf{C}}_j,{{\fancyscript{R}}}_J) +o(R_J^{\it v}) $$
(5.223)

where \(s=s_{\rm min}(n_\uparrow,n_\downarrow),\) while \(l>0\) and \(\varepsilon=\pm1\) are parameters of the model playing a role analogous to the absolute value and the sign of the two-body scattering length. This approach is only possible if the wavefunction remains square integrable, i.e. if \(0\le s<1\), which we assume in what follows. This condition is satisfied e.g. for \(n_\uparrow=2, n_\downarrow=1\) for a mass ratio \(m_\uparrow/m_\downarrow \in \, [8.62\ldots;13.6\ldots]\)[5]. Moreover we are assuming for simplicity that \(s\neq0.\)

Let us now consider the particular case where the two-body scattering length is infinite, and the external potential is either harmonic isotropic, or absent. Then the separability in internal hyperspherical coordinates of Sect. 5.3.3 still holds for \(n=N\). Indeed, Eq. 5.223 then translates into the boundary condition on the hyperradial wavefunction

$$ \exists A\in{\mathbb{R}}/\ F(R){\mathop = \limits_{{R \to 0}}} A \cdot \left( R^{-s} - \frac{\varepsilon}{l^{2s}} R^s \right) + O\left(R^{s+2}\right) $$
(5.224)

and does not affect the hyperangular problem. Consequently [55],

  • For the n-body bound state, which exists if \(\varepsilon=+1\):

    $$ E=-\frac{2\hbar^2}{m\,l^2} \left[\frac{\Upgamma(1+s)}{\Upgamma(1-s)}\right]^{\frac{1}{s}}, $$
    (5.225)
    $$ F(R)=K_s\left(R \sqrt{-2 E \frac{m}{\hbar^2}}\right). $$
    (5.226)
  • For the eigenstates in a trap:

    $$ E { \hbox{solves:} } -\varepsilon\cdot\left(\frac{\hbar}{m\omega\,l^2}\right)^s = \frac{\Upgamma\left( \frac{1+s-E/(\hbar\omega)}{2} \right) \Upgamma(-s)} {\Upgamma\left( \frac{1-s-E/(\hbar\omega)}{2} \right) \Upgamma(s)}, $$
    (5.227)
    $$ F(R)=\frac{1}{R}\,W_{\frac{E}{2\hbar\omega},\frac{s}{2}}\left(R^2\frac{m\omega}{\hbar}\right). $$
    (5.228)

In particular, for \(l=\infty\), we are exactly at the n-body resonance, since the energy of the n-body bound state vanishes. The spectrum in a trap then is \(E=(-s+1+2 q)\hbar\omega\) with \(q\in\mathbb{N}.\)

Note that, most often, \(s\,{\geq}\,1,\) in which case one would have to use an approach similar to the one developped by Pricoupenko for the case of two-body resonances in non-zero angular momentum channels, and to introduce a modified scalar product [22, 104].

Rights and permissions

Reprints and permissions

Copyright information

© 2012 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Castin, Y., Werner, F. (2012). The Unitary Gas and its Symmetry Properties. In: Zwerger, W. (eds) The BCS-BEC Crossover and the Unitary Fermi Gas. Lecture Notes in Physics, vol 836. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-21978-8_5

Download citation

  • DOI: https://doi.org/10.1007/978-3-642-21978-8_5

  • Published:

  • Publisher Name: Springer, Berlin, Heidelberg

  • Print ISBN: 978-3-642-21977-1

  • Online ISBN: 978-3-642-21978-8

  • eBook Packages: Physics and AstronomyPhysics and Astronomy (R0)

Publish with us

Policies and ethics