Bose Einstein Condensation Compress
Bose Einstein Condensation Compress
Edited by
A. GRIFFIN
University of Toronto, Toronto, Canada
D. W. SNOKE
University of Pittsburgh, Pittsburgh, USA
S. STRINGARI
University of Trento, Povo, Italy
CAMBRIDGE
UNIVERSITY PRESS
Published by the Press Syndicate of the University of Cambridge
The Pitt Building, Trumpington Street, Cambridge CB2 1RP
40 West 20th Street, New York, NY 10011-4211, USA
10 Stamford Road, Oakleigh, Melbourne 3166, Australia
TAG
Contents
Preface page xi
Preface to paperback edition xiii
VII
viii Contents
15 Crossover from BCS Theory to Bose-Einstein Condensation
M. Randeria 355
16 Bose-Einstein Condensation of Bipolarons in High-T c
Superconductors J. Ranninger 393
17 The Bosonization Method in Nuclear Physics F. Iachello 418
18 Kaon Condensation in Dense Matter G. E. Brown 438
19 Broken Gauge Symmetry in a Bose Condensate A. J. Leggett 452
XI
xii Preface
overdue. The idea then developed that the main focus of this first BEC
workshop should be review-type talks by carefully chosen international
experts and that these talks would be the basis of an authoritative book
on BEC research, suitable for someone just getting interested in this topic.
The International Workshop on Bose-Einstein Condensation (BEC 93)
was duly organized and held in Levico Terme, a small resort town near
Trento in northern Italy, 31 May to 4 June, 1993. It was a resounding
success from every point of view. Over 70 participants from over 20
countries (listed in an appendix at the end of this book) represented all
the major research groups currently working on BEC, and included a
good mix of experimentalists and theorists. Everyone had a sense that
this first BEC workshop was an historic occasion, helped by the fact
that the participants included many of the "founding fathers" of BEC
studies, such as K. Huang, S.A. Moskalenko and L.V. Keldysh. The
unifying themes of BEC research were stressed by having talks on quite
different sub-fields given each day.
We decided that the invited review articles could be submitted well
after the workshop, so that they would reflect the excitement and the
interdisciplinary insights which arose from the workshop. The present
volume contains a few additional invited review articles covering topics
which were not emphasized at BEC 93. At the workshop, over thirty
poster contributions were also presented. The second part of this book
contains brief research reports and mini-reviews based on some of these
contributions. The refereeing was done by the editors and the authors
of the invited papers. Many of the contributed papers have since been
submitted to research journals and for this reason are not included here.
There is only one review article devoted to superfluid 4He. It was felt
that since a recent book has given an extensive review of superfluid 4He
(A. Griffin, Excitations in a Bose-Condensed Liquid, Cambridge University
Press, 1993), it was not necessary to emphasize this subject in the present
volume.
The organizing committee of BEC 93 at Levico Terme consisted of
the editors, with the assistance of F. Laloe and G. Baym. In addition,
A. Mysyrowicz, S. Tikhodeev and D. van der Marel were very supportive
at the very early stages of organization. The major source of external
funding of BEC 93 was from the European Centre for Theoretical Studies
in Nuclear Physics and Related Areas (ECF), which has recently been
set up at the University of Trento. In addition, BEC 93 was financially
assisted by grants from CNRS (France) and the Department of Physics
at the University of Toronto, as well as funds from NSERC of Canada.
Preface xiii
A crucial role was played by the workshop secretary, Jennifer Tarn
(Toronto).
The local organization at the University of Trento and Levico Terme
was carried out by Sandro Stringari, and we thank especially F. Dalfovo
and S. Giorgini for their help. The Grand Hotel Bellavista in Levico
Terme played an important role in producing a successful workshop,
providing the right atmosphere as well as fine Italian cuisine. Late into
each evening, heated discussions on topics such as timescales and broken
symmetry could be heard coming from the hotel bar.
Preparation of the book in the Cambridge University Press L£F]gX
format was carried out by David Snoke and Karl Schroeder (Toronto).
Most of the contributors made this job much easier for us by sending
in their articles in TgX or L^TgX. We would also like to thank Simon
Capelin of Cambridge University Press for his enthusiastic support right
from the beginning.
We hope this unique volume will introduce many readers to the chal-
lenging problems involved in the quest for BEC and set the stage for
future progress.
Allan Griffin, Toronto
David Snoke, Pittsburgh
Sandro Stringari, Trento
Gordon Baym
Department of Physics
University of Illinois at Urbana-Champaign
1110W. Green St, Urbana, IL 61801, USA
References
[1] J. Gasser and H. Leutwyler, Phys. Reports 87, 77 (1982).
[2] C. T. Hill, Phys. Lett. B 266, 419 (1991) and references therein.
[3] See, e.g., International conferences on quark matter and ultra-
relativistic nucleus-nucleus collisions, Nucl. Phys. A 544 (1992); in
press (1994).
[4] P. C. Hohenberg and P. M. Platzman, Phys. Rev. 152, 198 (1966).
[5] D.M. Ceperley and E.L. Pollock, Phys. Rev. Lett. 56, 351 (1986).
[6] A. Griffin, Excitations in a Bose-Condensed Liquid (Cambridge Uni-
versity Press, Cambridge, 1993).
[7] For a review, see D.R. Nelson, in Phase Transitions and Critical
Phenomena 7, C. Domb and J.L. Lebowitz, eds. (Academic, New
York, 1983).
[8] M. Rho, "Cheshire Cat Hadrons," Phys. Reports, in press (1994).
[9] See, e.g., V.N. Popov, Functional Integrals and Collective Excitations,
(Cambridge University Press, Cambridge, 1987), ch. 6.
Introduction: Unifying Themes of Bose-Einstein Condensation 11
[10] N. Peghambarian, L.L. Chase, and A. Mysyrowicz, Phys. Rev. B 27,
2325 (1983).
[11] W. Kohn and D. Sherrington, Rev. Mod. Phys. 42, 1 (1970).
[12] E. Hanamura and H. Haug, Sol. State Comm. 15, 1567 (1974); Phys.
Reports 33, 209 (1977).
[13] KM. Peeters and J.E. Golub, Phys. Rev. B 43, 5159 (1991).
[14] T. Fukuzawa et a/., Phys. Rev. Lett. 64, 3066 (1990).
[15] J.A. Kash et al, Phys. Rev. Lett. 66, 2247 (1991).
[16] D.S. Chemla J.B. Stark, and W.H. Knox, Ultrafast Phenomena VIII,
(Springer, Berlin, 1993), p. 21.
[17] I.V. Lerner and Yu. E. Lozovik, Sov. Phys. JETP 53, 763 (1981).
[18] D. Paquet et al, Phys. Rev. B 32, 5208 (1985).
[19] The analogy between a laser and superfluid 4 He is nicely reviewed
by PC. Martin, in Low Temperature Physics— LT9, J.E. Daunt,
D.O. Edwards, F.J. Milford, and M. Yaqub, eds. (Plenum, New
York, 1965), p. 9.
[20] See, e.g., H. Haken, Rev. Mod. Phys. 47, 67 (1975).
[21] J. Madsen, Phys. Rev. Lett. 69, 571 (1992).
[22] N. Kaiser, R. A. Maleney, and G. D. Starkman, Phys. Rev. Letters
71, 1128 (1993).
[23] Finite size effects are discussed by E.L. Pollock, Phys. Rev. B 46,
3535 (1992).
[24] Ph. Sindzingre, M. Klein and D. M. Ceperley, Phys. Rev. Lett. 63,
1601 (1989).
Part one
Review Papers
Some Comments on Bose-Einstein
Condensation
P. Nozieres
Institut Laue-Langevin
B.P. 156, 38042 Grenoble Cedex 9
France
Abstract
The paper reviews some important features of Bose-Einstein condensation
such as the nature of the condensate, exchange interactions, phase locking.
The competition with other physical effects (crystallization, dissociation,
localization) is briefly discussed.
15
16 P. Nozieres
(1)
o
r^ ^ T
p i
c N
The fully condensed ground state has all particles in the lowest state
N
J5) |vac>. (2)
The corresponding interaction energy is
2
\v(b*b*bb)
Eo = \v ^VN(Nl)
o(b*ob*obobo) = ^VoN(N-l) ~ ±VN
±V0N
Note that Vo is necessarily > 0 (i.e., repulsive): otherwise the system
would collapse spontaneously to infinite densities JV.
We compare that state with one in which the condensate is shared
between states 1 and 2, with occupancies JVi and N2 = N — N\
IW = J^^^'^^lvac). (3)
The kinetic energies are both close to 0 and negligible. The interaction
energy involves all possible contractions of operators in the expectation
value of V. It contains a Hartree term (between all species) and a Fock
exchange term. The latter only occurs between different species, since
there is only one type of contraction in the expectation value of b\b\b\b\.
Hence
Hartree Fock
I 1
E = y [ATi(ATi - 1) + N2(N2 - 1) + 2iViJV2] + V^
(4)
(( ) (5)
a,a'
Here we neglect the depletion via virtual excitation which should also
depend on the internal state a. The selection argument relies only on
intracondensate interactions, which is certainly an approximation. We
note first that the energy £o must be positive for any choice of na > 0:
otherwise the density would collapse. The nature of the minimum then
depends on the matrix A. Consider, for instance, a two-fold situation
d8e-iNe\xp0(8)).
Jo
It is easily verified that (N) = |0| 2 : the energy is the same in (2) and (6).
The situation is different when we consider interaction terms such as
b^bkb-k + c.c.
which allow virtual excitations of two particles out of the condensate.
In order to take such a hybridization into account, we try a variational
wave function
\xp0) = (7)
Some Comments on Bose-Einstein Condensation 21
Afe is obtained minimizing the energy. Equation (7) is nothing but the well-
known Bogoliubov approximation for Bose liquids. When calculating the
expectation value of the energy, we recover the usual Hartree and Fock
exchange contributions, which here have corrections of order
V0\h\2 , Vk\Xk\2.
They are the same whether N or 9 is locked. The novelty is the so-called
Bogoliubov term, which appears only when the phase is locked
Such a linear coupling between <j> and kk always lowers the energy, by
an extensive amount of order V2: hybridization is bound to occur. It
follows that phase locking is implied by Bose-Einstein condensation. Note
that the Bogoliubov energy favours pure state condensates: it is enough
to lift the degeneracy.
Phase locking is a genuine symmetry breaking, distinct from Bose-
Einstein condensation. It is responsible for superfluidity (superfluid
currents are due to a gradient in the locked phase 6). Its physical
consequences are well known : quantization of circulation in multiples
of h/m, vortices, Josephson effect through a weak link, etc.
As an unconventional example let us consider Bose condensation of
excitons in CU2O, extensively discussed by Wolfe et al. and Mysyrow-
icz [3]. These excitons exist in two brands, para (lowest in energy) and
ortho (some 12 meV higher). They are seen through phonon assisted
fluorescence (momentum k is not conserved), from which one infers the
exciton energy distribution n(e). Most of the data are concerned with
orthoexcitons, less "forbidden" than the para. As the density N grows,
quantum deviations to n(e) are clearly visible, but somehow one never
reaches Bose-Einstein condensation. One "surfs" on the transition line
without being able to cross it - why?
A tentative explanation might go as follows. Being lower, the paraex-
citons should be more numerous and consequently they should Bose
condense first. But they are harder to see: their condensation may es-
cape observation. Assume now that the ortho density grows through the
transition. Then a Josephson process is possible
O + O-+P +P.
Direct conversion of a single exciton is forbidden by symmetry, but
there is no such problem for a pair. That Josephson current is a.c, at
a frequency co corresponding to the 12 meV splitting - hence a strong
22 P. Nozieres
dissipation that precludes further cooling. One thereby would explain
why any onset of ortho condensation would immediately stop the process.
The above example may be wrong, but it points to an essential fact:
phase locking is a crucial feature of condensed systems.
4.1 Crystallization
The energy per particle of a liquid phase is roughly an interaction Vm
averaged over a unit cell (the distance between atoms is free to vary,
except for some short range correlations). This may be compared with a
solid phase in which the particles are localized at lattice sites. The latter
energy contains a kinetic part due to localization of the particles, and a
potential energy Va corresponding to the lattice spacing a
N ma2
Because the repulsive potential V(r) decreases with r, we have Va < Vm.
It follows that the crystalline state will be favoured for strong coupling (it
optimizes the potential energy), while the Bose condensed liquid will win
at weak coupling (it minimizes the kinetic energy). An obvious example
is 4He, which is a superfluid at low pressure and which freezes at higher
densities.
Note that the kinetic energy that is relevant to the above competition
is that associated to coherent hopping of the particle from place to
place, from which one may construct Bloch states (remember that phase
coherence is an essential ingredient). Such a coherence may be very hard
to achieve. Consider for instance bipolarons, i.e. two fermions trapped in
the same cloud of phonons that they have generated. These bipolarons
are candidates for high temperature superconductivity. What matters
here is elastic coherent hopping of the bipolaron, leaving the phonon
bath in its ground state. The corresponding amplitude W involves the
overlap of phonon ground states O, and O7 when the particle is at sites i
or;,
W =
Some Comments on Bose-Einstein Condensation 23
This overlap is usually very small and the kinetic energy is negligible
as compared to potential energy. Charge density waves (crystallization)
win!
The above competition is nicely formulated in the old lattice gas model
of Matsubara and Matsuda [4]. Consider N hard core bosons on 2N
sites. Due to the single occupancy constraint, each site has two states,
empty or filled. The bosons have commutation rules on different sites,
which is just the algebra of spins. The hard core boson gas is therefore
isomorphous to a spin 1/2 system. The hopping amplitude between
nearest neighbours is equivalent to a transverse (xy) magnetic coupling:
q is the total momentum of the pair, cpk its internal wave function. The
Hartree ground state for the pairs (equivalent to (6)) is just
(8) is nothing but the usual BCS wave function, viewed here as pair
Bose-Einstein condensation. Vk is a variational parameter which allows
an interpolation between low and high densities. (Note that (8) ignores
the effect of screening and its interplay with binding: the real problem of
Mott transitions is not tackled.)
The superfluid state (8) exists only if the particles are subject to an
attraction. Two limiting cases are simple:
(i) In the dilute "molecular" limit the attraction must be strong enough
to create a bound state of two fermions ("dilute" means a particle
separation large as compared to the bound state radius). The bound
pairs behave as "point" bosons that condense in the usual way. Vk is
<C 1 and it reflects the internal state of the underlying fermions. The
gap A is the molecular binding energy, while the critical temperature
Tc is controlled by centre of mass motion of the bound pairs. For free
bosons Tc ~ AT5, while for a Hubbard lattice gas Tc « 1/17 (one must
virtually break the pair in order to make it hop).
(ii) In the opposite dense limit, the pairs strongly overlap and the exclusion
principle becomes the dominant feature. \vk\2 cannot exceed 1: it
extends to higher k in order to accommodate the N particles - in
practice to the Fermi momentum kF. In the limit of high densities
we recover the usual BCS result, with an exponentially small gap A.
The fermions are normal except for a narrow region of width A near
Some Comments on Bose-Einstein Condensation 25
Fig. 4. The evolution of the BCS parameter x>\ as a function of energy efc for
increasing densities. Note the saturation due to the exclusion principle.
Strictly speaking, the effect does not depend on statistics - except that
something must prevent accumulation of particles in the lowest local-
ized states. For fermions the exclusion principle does it. For bosons
one must rely on repulsive interactions that forbid large local densi-
ties. At first sight one expects a transition from a localized state at
small coupling - the so-called Bose glass - to a regular superfluid at
large coupling, when interactions force the chemical potential \i into
the region of extended states. But one might also argue that strong
Some Comments on Bose-Einstein Condensation 27
(ii) Interactions between particles are absolutely crucial - one may say
that genuine condensation is an effect of exchange coupling.
References
[1] V. Bagnato and D. Kleppner Phys. Rev. A 44, 7439 (1991).
[2] S.I. Sevchenko, Sov. Phys. JETP 73, 1009 (1991).
30 P. Nozieres
[3] J.R Wolfe et al. and A. Mysyrowicz, this volume.
[4] T. Matsubara and H. Matsuda, Prog. Theoret. Phys. 16, 569 (1956);
17, 19 (1957).
[5] K.S. Liu and M.E. Fisher, J. Low Temp. Phys. 10, 655 (1973).
[6] A.F. Andreev and I.M. Lifshitz, Sov. Phys. JETP 29, 1107 (1969).
[7] G.A. Lengua and J.M. Goodkind, J. Low Temp. Phys. 79, 251 (1990).
[8] R. Ma, B.I. Halperin and PA. Lee, Phys. Rev. B 34, 3136 (1986);
M.P.A. Fisher, RB. Weichman, G. Grinstein and D.S. Fisher, Phys.
Rev. B 40, 546 (1989). An excellent review by TV. Ramakrishnan is
in press.
[9] N. Trivedi, in Computer Simulation Studies in Condensed Matter
Physics V, (Springer, Berlin, 1993); W. Krauth, N. Trivedi and D.
M. Ceperley, Phys. Rev. Lett. 67, 2307 (1991).
[10] D.B. Haviland, Y. Liu and A.M. Goldman, Phys. Rev. Lett. 62, 2180
(1989).
Bose-Einstein Condensation and
Superfluidity
Kerson Huang
Department of Physics
and
Center for Theoretical Physics, Laboratory for Nuclear Science
Massachusetts Institute of Technology
Cambridge, MA 02139
USA
Abstract
We review generally accepted definitions of Bose-Einstein condensation and
superfluidity, emphasizing that they are independent concepts. These ideas
are illustrated in a dilute hard-sphere Bose gas, which is relevant to ex-
periments on excitons and spin-aligned atomic hydrogen. We then discuss
superfluid 4He in porous media, as simulated by different models in differ-
ent regimes. At low coverage, we model it by a dilute hard-sphere Bose gas
in random potentials, and show that superfluidity is destroyed through the
pinning of the Bose condensate by the external potentials. At full coverage,
we model the random medium by an ohmic network of random resistors,
and argue that the superfluid transition is a percolation transition in d=3,
with critical exponent 1.7.
1 Bose-Einstein Condensation
In an ideal Bose gas in three spatial dimensions (d=3), Bose-Einstein
condensation occurs when there is a finite density of particles in the
31
32 K. Huang
zero-momentum (k=0) state:
(alao)>O (1)
f ~ e^\a\ak) (2)
= N (5)
(A}Ai) = N/K
which lead to the relation (ajao) = K"1 ^(AJAi)+K~l J2i+j(AlAj)> or
N
= £ + £ H(AUj) (6)
For large K, the second term on the right side must dominate, and
therefore the phases of the At in different boxes must be correlated.
The correct order parameter for Bose-Einstein condensation is thus
the condensate wave function. As defined, however, it is a micro-
scopic property of the system, and we can rarely make headway with
microscopic calculations. It is more practical to adopt the Ginsberg-
Landau approach, and introduce a local order parameter </>(r) that is a
coarse-grained version of (y>(r)):
</>(r) = jR{r)em (7)
whose phase is related to the superfluid velocity through
ys = - V a (8)
m
where m is the boson mass. Since 0 should be single-valued, the change
of a over any closed path must be a multiple of 2n. This leads to the
Onsager-Feynman quantization condition [10, 11]
0c = 0,±l,±2,---) (9)
m
The Landau free energy is taken to be [12]
(d=3)
where b is a constant. We see that phase coherence is maintained to
infinite distances in d=3, but not in d=2 at finite temperatures. A
rigorous proof of the absence of Bose-Einstein condensation in d=2 at
finite temperatures was given by Hohenberg [18]. Equivalent statements
are that long-ranged order does not exist in d=2 (Mermin-Wagner
theorem) [19], and that spontaneous symmetry breaking does not occur
in d=2 (Coleman's theorem) [20].
Even though the correlation function in d=2 goes to zero asymp-
totically, it does so slowly, with power-law behavior. This means that
phase coherence persists over a finite but large distance, and a local Bose
condensate exists. The local Bose condensate supports vortex-antivortex
Bose-Einstein Condensation and Superfluidity 35
pairs, whose unbinding leads to the Kosterlitz-Thouless transition [21].
Thus, despite the fact that there is no Bose condensate, there is a phase
transition, associated with what Kosterlitz and Thouless termed "topo-
logical order."
In the Ginsberg-Landau picture, the Kosterlitz-Thouless transition in
d=2 should not be associated with a sign change of c^. It is caused
by vorticity associated with phase fluctuations, and to support vortices
we must must have ci < 0. It would be interesting to discover such
a transition within the Ginsberg-Landau theory; but this has not been
attempted. Instead, one takes the two-dimensional XY model [22] as
a more expedient starting point. As we shall discuss later, this model
exhibits superfluidity in the low-temperature phase, even though there is
no Bose condensate.
The local Bose condensate in d=2 is relevant to current thinking in
high-Tc superconductivity, where one speaks of Bose-Einstein condensa-
tion of "holons" in the two-dimensional copper oxide plane. What one
means is a "local condensation", which presumably becomes global when
a small amount of interplane coupling is turned on.
2 Superfluidity
The term superfluidity covers a group of experimental phenomena in liq-
uid 4He, including frictionless flow, persistent current, wave propagation
on liquid surfaces, and other kinetic effects. Our understanding on this
subject is considerably less than that for Bose-Einstein condensation.
These first attempts to define superfluidity were based on specific
pictures of liquid 4He. Landau [23] proposed a low-temperature model
in which the excitations are quasiparticles, which become phonons in
the long-wavelength limit. Feynman [24] justified the picture from a
microscopic point of view, and argued that Bose-Einstein statistics rule
out all excitations except density fluctuations (phonons) in the immediate
neighborhood of the ground state.
Landau argued that a system with only phonon excitations will flow
without friction, because it cannot absorb arbitrarily small amounts of
energy-momentum transfer. The argument is as follows. At absolute zero
the only way to interact with the system is to excite phonons. Suppose
an external object transfers energy AE and momentum AP by creating
np phonons of momentum p and energy cp, where c is the sound velocity.
36 K. Huang
Then
AE = Y^ cpnp
p
V (14)
Thus |AP| < Y,PPnP> o r AE > c|AP|. If the external object is moving
with velocity ye, we must have AE = ye • AP. Therefore ve > c.
Using the quasiparticle picture, Khalatnikov [25] derived an expression
for the normal fluid density at finite temperatures, based on Gallilean
invariance. Suppose the energy of a quasiparticle of momentum p is
(Op, and the average occupation number for quasiparticles is n(co) =
[exp(co//cT) — I]" 1 . In a two-fluid picture, with superfluid velocity vs and
normal fluid velocity vn, the quasiparticle energy in a reference frame
moving with the superfluid is
(o'p = (op-p'(yn-ys). (15)
The current density in the co-moving frame is given by
= pn(yn-ys) (16)
where the right side is a definition of the normal fluid mass density pn.
By expanding in powers of (vn — vs) and identifying coefficients on both
sides, one easily obtains
Pn
~ 3j (2nfP dop (17)
6v
5v
(a) (b)
Fig. 1. (a) Longitudinal response: Place liquid between parallel plates that are
moving at velocity S\. The response gives the total density, (b) Transverse
response: Place liquid in long pipe moving with velocity S\. The response gives
the normal fluid density.
which defines the susceptibility %i;(k,co). We are concerned only with the
static susceptibility &/(k) = #i 7(k,0). By rotational invariance, it can be
decomposed into a longitudinal and a transverse part:
In the limit k —> 0, the longitudinal response is the total density (statement
of the f-sum-rule), while the transverse response is defined as the normal
fluid density. Thus, we have
p = lim A(k2)
Transition line
-• V
are
a/k
na3 (23)
k= ^2nh2/mkT (24)
F(r) = (25)
m
Using this, we can get Lenz's result in first-order perturbation theory.
40 K. Huang
With the pseudopotential, the many-body Hamiltonian reads
(r)xp(r)xp{r)
44 ' (26)
k,p,q J
where Q is the volume of the system. The ground state energy is divergent
in higher orders; but it can be made finite by a simple subtraction
procedure [36].
The Hamiltonian can be diagonalized using Bogoliubov's method [37],
in which ao is replaced by a c-number y/No. The parameter No, which
labels the state we are considering, is the number of fc=0 particles in the
unperturbed state. We write it in the form
No = £N, (27)
where N is the total number of particles. The diagonalized form of the
Hamiltonian to order a5/2 is given by [38, 39]
H =
(28)
0k = / (fok — ^k^k)
where
. (30)
This shows that the interactions deplete the Bose condensate; but the
superfluidity density at absolute zero is n, from Khalatnikov's formula
(17).
The partition sum should extend over all values of £. For thermal
properties near the transition point, the important states have £ « 0, and
in first approximation we keep £ only to lowest order. This means that
the quasiparticle spectrum is replaced by a shifted free-particle spectrum
(ft2/2m)(k2 + Snan£). The calculation of the free energy is elementary [39],
and gives
(( h ) (32)
m \ 2
where F^ is the free energy of the ideal Bose gas, and | , the thermody-
namic average of £, is the same as that in the ideal Bose gas:
j_jl-{T/Tcf'2 (T<TC)
1.0 (T>TC),
(a) (b)
Fig. 4. (a) Transition surface for superfluid transition of 4He in porous media.
Superfluidity disappears at T=0 below a critical coverage. After Physics Today,
July, 1989, p.22. (b) Transition surface for superfluid transition in a hard-sphere
Bose gas in random external potentials in d=3. The total density is n, and the
superfluid particle density is ns.
n = n 0 + m + n R, (37)
where n\ arises from the hard-sphere interactions, and can be read off
(31), while nR arises from the random potentials:
Bose-Einstein Condensation and Superfluidity 45
ni = .
\y%t\
(38)
-I
p(E) = / d\ d(E - ck) oc E2/c\ (41)
The remarkable factor 4/3 means that the random potentials generate
an amount of normal fluid 4/3 times what it took from the condensate.
This implies that part of the condensate belonging to the normal fluid,
being pinned by the random scatterers - a example of boson localization.
We can now see that superfluidity can be destroyed, even though there
46 K. Huang
is a Bose condensate. From (37) and (42), we can deduce the relation
n, = ^[4(no+ * ! ) - * ] , (43)
-5 H Vycor(2.00)
* Xerogel ( 2.09 )
, I I . . .
-5 -4 -3 -2 -1
Log(T c -T)
Fig. 5. Log-log plot of superfluid density vs. temperature for liquid 4He in fully
saturated media. The number in braces after the name of each medium gives
the value of Tc used in the plot. Percentages correspond to porosity of aerogel
samples [51].
flow is given by
j = PsVs + Pn (44)
where the two terms refer to the superfluid and normal fluid respectively.
Assuming that all channels are so small that the normal fluid is pinned,
we put \n = 0. Thus, writing vs = V®, where O is proportional to the
superfluid phase, we have
j = PsV<D (45)
This looks like Ohm's law with current density j , electrostatic potential
O and electrical conductivity ps. Thus, each bond can be replaced by
a resistor, and the superfluid density of the medium corresponds to the
conductivity of the network.
In a long channel filled with liquid 4He, the superfluid transition
temperature decreases with the radius according to a power law [52]. At
sufficiently high temperatures, all bonds in the lattice are non-conducting,
and the overall conductance is zero. As the temperature is lowered, some
bonds begins to conduct; but the overall conductance remains zero,
until there is a percolating cluster. Therefore, the true critical point
corresponds to percolation in d=3.
This picture is verified by solving Kirchoff's equations numerically to
obtain the conductivity of the network, for various distributions of the
48 K. Huang
resistances. The critical exponent is independent of the distribution used.
However, the true critical region is very narrow. At lower temperatures,
there is a wide "mean-field" region in which the numerically calculated
superfluid density can be fit with a power law, whose exponent depends
on the distribution. By choosing the distribution appropriately, we can
reproduce the apparent exponent observed for the different media. In
Fig. 5 we plot all available data on a log-log plot to show that they do
not contradict our picture.
This work is supported in part by funds provided by the US Depart-
ment of Energy under contract # DE-AC02-76ER03069.
References
[I] S.N. Bose , Z. Phys. 26, 178 (1924).
[2] A. Einstein, Sitzber. Kgl. Preuss. Akad. Wiss., (1924), p.261 ; (1925),
p.3 . The story of Bose-Einstein statistics is told in A. Pais, Subtle
is the Lord, The Science and the Life of Albert Einstein (Clarendon
Press, Oxford, 1982), Ch. 23.
[3] L. Tisza, Nature, 141, 913 (1938),
[4] O. Penrose and L. Onsager, Phys. Rev. 104, 576 (1956).
[5] RE. Sokol, this volume.
[6] J. Goldstone, N. Cim. 19, 154 (1961).
[7] C.N. Yang, Rev. Mod. Phys. 34, 4 (1962).
[8] RW. Anderson, Rev. Mod. Phys. 38, 298 (1966).
[9] RW. Anderson, Basic Notions of Condensed Matter Physics
(Benjamin-Cummings, Menlo Park,1984), pp.l5ff.
[10] L. Onsager, Suppl. N. Cim. 6, 249 (1949).
[II] R.P. Feynman, in Progress in Low Temperature Physics, vol.1, C.J.
Gorter, ed. (North-Holland, Amsterdam, 1955), p. 17.
[12] V.L. Ginsberg and L.R Pitaevsky, Sov. Phys. JETP 34, 858 (1958).
[13] K.G. Wilson and J. Kogut, Phys. Rep. 12C, 75 (1974).
[14] E.P. Gross, N. Cim. 20, 454 (1961); J. Math. Phys. 4, 195 (1963).
[15] L.R Pitaevsky, Sov. Phys. JETP 13, 451 (1961).
[16] T. Frisch, Y. Pomeau, and S. Rica, Phys. Rev. Lett. 69, 1644 (1992).
[17] See the contributions of Yu. Kagan and H.T.C. Stoof in this volume.
[18] RC. Hohenberg, Phys. Rev. 158, 383 (1967).
[19] N.D. Mermin and H. Wagner, Phys. Rev. Lett. 22, 1133 (1966).
[20] S. Coleman, Comm. Math. Phys. 31, 259 (1973).
[21] J.M. Kosterlitz and D.J. Thouless, J. Phys. C 6, 1181 (1973).
Bose-Einstein Condensation and Superfluidity 49
[22] J.V. Jose, L.R Kadanoff, S. Kirkpatrick, and D.R. Nelson, Phys.
Rev. B 16, 1217 (1977).
[23] L.D. Landau, J. Phys. USSR 5, 71 (1941).
[24] R.P. Feynman, Phys. Rev., 94, 262 (1954).
[25] I.M. Khalatnikov, Introduction to the Theory of Superfluidity (Ben-
jamin, New York, 1965), p.13.
[26] PC. Hohenberg and PC. Martin, Ann. Physics (NY) 34, 291 (1965).
[27] G. Baym, in Mathematical Methods of Solid State and Super fluid The-
ory, ed. R.C. Clark (Oliver and Boyd, Edinburgh, 1969), p.121; the
argument is reproduced in D. Forster, Hydrodynamics Fluctuations,
Broken Symmetry, and Correlation Functions (Benjamin, Reading,
MA, 1975), pp.219 ff.
[28] D.R. Nelson and J.M. Kosterlitz, Phys. Rev. Lett. 39, 1201 (1977);
D.R. Nelson in Phase Transitions and Critical Phenomena, vol.7, C.
Domb and J.L. Lebowitz, eds. (Academic, New York, 1983).
[29] I. Rudnick, Phys. Rev. Lett. 40, 1454 (1978); D J . Bishop and J.
Reppy, Phys. Rev. Lett. 40, 1727 (1978); E. Webster, G. Webster,
and M. Chester, Phys. Rev. Lett. 42, 243 (1978); J.A. Roth, G.J.
Jelatis, and J.D. Maynard, Phys. Rev. Lett. 44, 333 (1978).
[30] See the article by L. V. Keldysh, this volume.
[31] See the article by A. Leggett, this volume.
[32] See the articles by T. Greytak and I. Silvera, this volume.
[33] See the article by J. P. Wolfe, et al, this volume.
[34] W. Lenz, Z. Phys. 56, 778 (1929).
[35] E. Fermi, Riverca Sci. 7, 13 (1936).
[36] K. Huang and C.N. Yang, Phys. Rev. 105, 767 (1956).
[37] N.N. Bogoliubov, J. Phys. USSR, 2, 23 (1947).
[38] T.D. Lee, K. Huang, and C.N. Yang, Phys. Rev. 106, 1135 (1957).
[39] K. Huang, in Studies in Statistical Mechanics, vol.2, J. De Boer and
G.E. Uhlenbeck, eds. (North-Holland, Amsterdam, 1964).
[40] K. Huang, C.N. Yang, and J.M. Luttinger, Phys. Rev. 105, 776
(1957).
[41] K. Huang, Phys. Rev. 115, 765 (1959); 119, 1129 (1960).
[42] J.A. Hertz, L. Fleishman, and PW. Anderson, Phys. Rev. Lett. 43,
942 (1979).
[43] D.S. Fisher and M.P.A. Fisher, Phys. Rev. 61, 1847 (1988).
[44] M.P.A. Fisher, P.B. Weichman, G. Grinstein, and D.S. Fisher, Phys.
Rev. B 40, 546 (1989).
[45] For a comprehensive review, see J.D. Reppy, J. Low Temp. Phys.
87, 205 (1992).
50 K. Huang
[46] J.D. Reppy, private communication.
[47] K. Huang and H.F. Meng, Phys. Rev. Lett. 69, 644 (1992).
[48] H.F. Meng, Superfluidity and Random Media, Ph.D. Thesis, Math-
ematics Department, MIT, 1993.
[49] H.F. Meng, Quantum Theory of Two-Dimensional Interacting Bo-
son Systems, (MIT Mathematics Department preprint, 1993).
[50] K. Huang and H.F. Meng, Phys. Rev. B 48, 6687 (1993).
[51] The data in Fig. 5 is from J. D. Reppy and H. M. W. Chan, private
communication.
[52] R. Donnelly, R. Hills, and P. Roberts, Phys. Rev. Lett. 42, 75 (1979).
Bose-Einstein Condensation in Liquid
Helium
P. E. Sokol
Physics Department
Pennsylvania State University
University Park, PA 16802
USA
Abstract
Liquid helium is the prototypical example of a superfluid - a liquid that
flows without viscosity and transfers heat without a temperature gradient.
These properties are intimately related to the Bose condensation that oc-
curs in this strongly interacting liquid. Bose condensation is most directly
observed in the single particle atomic momentum distribution, where the
Bose condensate appears as a delta function singularity. In this article, we
discuss the experimental techniques used to observe the condensate and the
current status of measurments of the Bose condensate in liquid helium.
1 Introduction
51
52 P. Sokol
interactions, all the atoms in the liquid would occupy a single momentum
state at zero temperature.
Liquid helium is also unique among the Bose systems considered in this
volume. Other systems, such as spin polarized hydrogen and excitons,
are expected to exhibit Bose condensation. (See the review articles by
Greytak and Silvera, Castin et al, Mysyrowicz and Wolfe et al. in this
volume.) At present, however, liquid helium is the only system where the
existence of an experimentally attainable Bose condensed phase is almost
universally accepted.
Unlike many of the other systems considered in this volume, liquid
helium is a strongly interacting system. The interaction between the
atoms, which is dominated by the hard core repulsion, is strong enough
that the properties of the liquid cannot be treated as a simple perturbation
of the Ideal Bose Gas, as is the case for more weakly interacting systems.
As we will see, these strong interactions have greatly complicated both
the theoretical and experimental studies of the Bose-condensed phase of
superfluid 4He.
We begin this review with a discussion of the Ideal Bose Gas, a model
system consisting of non-interacting particles. This model, while not
applicable to the strongly interacting liquid phase of helium, is useful in
that it provides a system where the effects of Bose condensation are clearly
delineated. The microscopic properties of liquid helium, which we discuss
next, are quite different from the ideal gas due to the strong interactions
between atoms. Despite these differences, similarities between the ideal
gas and the liquid systems remain. In particular, a Bose-condensed phase
still appears in liquid 4 He at low temperatures, although the fraction of
particles in the condensate is considerably reduced. Therefore, in this
review, the analogies between an ideal Bose gas and superfluid 4He are
stressed, even though the former does not exhibit the long-range phase
coherence associated with Bose-broken symmetry. (For a discussion of
this, see especially the articles in this book by Nozieres, Huang, Stringari
and Stoof.)
The single-particle momentum distribution is one of the few quantities
which directly reflects the appearance of a Bose condensate. Therefore,
we next turn our attention to the experimental technique that provides
information on the momentum distribution: Deep Inelastic Neutron
Scattering (DINS). This technique, due to the information on n(p) that it
provides, provides one of the few opportunities to determine no directly [4,
5,6].
Early attempts to measure the momentum distribution, and to obtain
BEC in Liquid Helium 53
direct information on the condensate using DINS, were limited by the
fluxes and energies of neutrons available. Recent advances in neutron
sources, which now have both higher fluxes and higher energies, have led
to a new generation of experiments. These new experiments are reviewed,
and the information that can be extracted, both directly and with the
help of models, are described.
Finally, we turn our attention to the direct evidence for the condensate
in liquid helium. The experimentally observed scattering shows distinct
changes consistent with the appearance of a condensate. However, the
sought after direct evidence, an observation of the (5-function singularity
associated with the condensate, is not observed. Nevertheless, comparison
to current theoretical predictions and to empirical expressions based on
realistic models give a strong case for a finite value of no and reasonable
results for its temperature and density dependence.
(a)
(b)
— Ideal Bose gas
- - - - Ideal classical gas
1.5
\
\ T = 50 K
(d)u
\ T= 3.14K
0.5
p(A"')
(d ) 1 000
T=0K
(d)u
T=2.5K
T=3.14K
400
0
V
Fig. 1. Temperature dependence of the momentum distribution in the Ideal
Bose Gas (solid line) and the classical result (dashed line). The Bose-Einstein
condensation temperature is 3.14 K, corresponding to the density of liquid helium
at SVP. (a) T = 50 K, (b) T = 6 K, (c) T = 2.5 K (a condensate is present but
is not visible in this diagram due to the singular behavior of the uncondensed
component) and (d) T = 0.
We now turn our attention to liquid 4He. Unlike the IBG, helium atoms
do interact with each other. Due to these interactions, the properties of
liquid helium are significantly different from those of the IBG. However,
one can still have a macroscopic number of atoms with zero momentum
(the Bose condensate). The interactions do, however, significantly reduce
the occupation of the ground state below its value in an ideal gas.
Liquid helium is, in one sense, a weakly interacting system. The
attractive part of the helium potential is quite weak [8], which, when
coupled with the light mass of 4 He, leads to a large zero-point motion
that prevents the atoms from being strongly localized even in the con-
densed phases. Thus, the liquid phase is not even stable until very low
temperatures and a solid phase, in which the atoms must be localized
about lattice sites, does not exist unless substantial external pressure is
applied [9]. This weak localization leads to a large overlap of the atomic
wavefunctions which make quantum statistics an important factor in
determining the properties of the condensed phases.
At the same time, liquid 4 He is also a very strongly interacting sys-
tem [10]. The hard core repulsion between atoms is very large due to the
closed shell electronic structure. The large zero-point motion would lead
to a large overlap of the core wavefunctions. The hard core repulsion
prevents this and leads to large repulsive interaction energies between
the atoms.
The effects of the interactions in liquid helium can be highlighted by
comparing the phase diagram of the IBG and liquid helium, as shown
in Fig. 2. The IBG exhibits a single-phase transition from the high
temperature gas phase to the Bose-condensed phase. This transition,
which is marked by the appearance of a finite fraction of particles in
the ground state, increases with increasing density. Liquid helium, on
the other hand, has a considerably more complex phase diagram [11].
There is no low density phase at temperatures where Bose condensation
might occur. At low temperatures, the low density gas is not stable and
condenses to form a high density liquid. Thus, there is no region where
Bose condensation of 4He atoms is possible where the interactions might
be considered weak.
Another difference between the liquid and the ideal gas is the depen-
dence of the transition temperature on density. In the IBG the transition
temperature increases with density, reflecting the fact that overlap of the
wavefunctions occurs at higher temperatures as the density increases. In
BEC in Liquid Helium 57
0.280
0.240 -
Solid
0.200 - -
u
0.120
A
0.080 -
0.040 -
1 naccessible
0.000
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0
T(K)
Fig. 2. The phase diagram of liquid helium in the temperature-density plane. The
regions marked inaccessible are not stable in the bulk liquid. The Bose-Einstein
transition temperature as a function of density is also shown.
n\p) = J (7)
Typical behavior for both n(p) and p\(r) for an interacting Bose fluid
shown in Fig. 3. The dynamical short-range correlation due to the hard
core repulsion governs the small-r behavior of p\{r) and the large-/? be-
havior of n{p). The effects of statistics play little role in this small-r range
and pi(r) would have a similar shape for Fermion or Boltzmann (class-
ical) particles. The effects of statistical correlations are most apparent in
Pi(r) at large r and in n(p) at small p. The condensate, which gives rise to
a finite value of pi(r) at infinity, manifests itself as a ^-function at p = 0
in nip).
A simple physical interpretation of pi(r) exists in the case of bosons
[15]. As can be seen from the definition of pi(r), it is the product of a
BEC in Liquid Helium 59
^ 0.5 -
0.150
0.125 —
0.100 —
0.075 —
0.050 —
0.025 —
0.000
(A"1)
Fig. 4. Theoretical calculations of the momentum distribution. The ground
state momentum distributions have been calculated using GFMC (solid) and
variational (dotted) techniques. The momentum distribution in the normal liquid
(dashed) has been calculated using PIMC techniques.
Fig. 4 shows two recent calculations of n(p) in the ground state, based
on GFMC [21] and variational [20] techniques.
The remainder of the momentum distribution, excluding the conden-
sate d -function, is known as the uncondensed momentum distribution
and exhibits several interesting features. At small p, which corresponds
to long-range interactions among the atoms, the effects of statistics are
the dominant factor. For example, singular behavior [18, 19] due to
the coupling of the condensate to long-wavelength collective excitations
(phonons) appears. This singular behavior can be obtained exactly at
small p, where the phonons are well defined, and goes as 1/p2 in the
ballistic regime and 1/p in the hydrodynamic regime. The variational
n(p) in Fig. 4 shows this behavior explicitly at small p. This singular
behavior is not present in the GFMC results, presumably due to the
relatively small size of the samples used in the numerical calculations.
Alternately, the large p behavior of n(p) is determined primarily by the
62 P. Sokol
short-range repulsive interaction between atoms, and the Bose statistics
play little role.
The great majority of numerical studies of the liquid have concentrated
on the ground state due to the intrinsic limitations of the GFMC and
variational approaches. However, Path Integral Monte Carlo (PIMC) [25]
methods have recently been applied to study the liquid properties at
finite temperatures. These calculations yield results similar to the ground
state calculations at low temperatures, i.e. a condensate fraction of
approximately 9%. However, they have the distinct advantage that they
can provide results at finite temperature. The condensate fraction is found
to exhibit a rapid increase of no entering the superfluid as temperature is
lowered, with very little variation with temperature thereafter.
The momentum distribution may also be calculated in the normal
liquid using PIMC. For example, Fig. 4 shows n(p) at 3.33 K, well above
the superfluid transition [25]. The momentum distribution is broad and
featureless with a nearly Gaussian form, the familiar classical result. The
width of the momentum distribution is determined by the quantum zero-
point motion of the liquid and is much larger than the width expected for
classical particles. However, aside from this, the shape of the momentum
distribution in the normal liquid shows little effect due to quantum
statistics, in contrast to the IBG discussed in Section 2.
KB-££« ft " ) + £
where aT, G\ and oc are the total, incoherent and coherent scattering cross
sections and kt and kf are the initial and final neutron wave vectors. The
interaction of the neutron with the sample is described by the coherent
and incoherent dynamic structure factors Sc(Q,co) and Si(Q,co), where co
and Q are the energy and momentum transfer of the scattered neutron.
The dynamic structure factors contain information on spatial and
temporal correlations between atoms in the sample [27]. The coherent
dynamical structure factor involves correlations between different atoms
BEC in Liquid Helium 63
and provides information on the structure and time-dependent behavior
related to phonons and rotons. The incoherent cross section, in contrast,
contains information only on the correlations of a single atom at two
different times.
The type of information provided by neutron scattering measurements
depends on the wavelength of the probing particle. Neutrons with wave-
lengths 2n/Q comparable to the interparticle separation provide infor-
mation on both the collective and single-particle correlations. However,
when the wavelength is much less than the average interparticle spacing,
neutron scattering measurements only probe single-particle correlations.
In this limit the wavelength of the neutron is too short to probe collec-
tive effects. The scattering is then determined totally by the incoherent
structure factor and provides information only on the single-particle mo-
tion of atoms. Even in this single-particle limit, however, the scattering
still probes the full many body properties of the sample through the
interactions of the scattering atom with the other atoms in the sample.
The scattering simplifies even further if the energy transferred to an
atom by a neutron is large compared with the interaction energies with
neighboring atoms. This limit [4], known as the Impulse Approximation
(IA), assumes that the neutron scatters from a single atom in a time
so short that the neighboring particles cannot react to the perturbation
caused by the neutron. The response of the He atom is then determined
entirely by n(p). It recoils from the collision in a free particle state of
high momentum and energy and the scattering directly probes the initial
many body wavefunction of the system.
The scattering in this high momentum and energy transfer limit is
described by the Compton profile
1 f+cc
JIA(Y) = -— 2 / pn(p)dp, (9)
4pnz J\Y\
where p is the density. The scattering in the IA does not depend on the
energy and momentum transfer separately, but only through the scaling
variable
Y=(M/Q)(co-corl (10)
where a> and Q are the energy and momentum transfer of the scattered
neutron, M is the mass and cor = Q2/2M is the recoil energy of the
scattering atom.
The scattering exhibits some characteristic features [28]; it is symmet-
ric, centered at Y = 0 and depends on a> and Q only through the scaling
64 P. Sokol
variable Y. These conditions have often been taken as indicative of
scattering in the IA limit. However, it is important to note that they are
necessary, but not sufficient, conditions for the validity of the IA. It is
possible for the scattering to exhibit these features even when the IA is
not valid, such as in the case of a gas of hard spheres [29].
The scattering in the IA limit is directly related to n(p) and, in the
case of a liquid, measurements of J(Y) can be used to determine n(p)
directly. For example, a Gaussian n(p) implies a Gaussian J(Y) with the
same second moment. Furthermore, the condensate, which appears as a
three-dimensional (5-function in n(p), is a one-dimensional (5-function in
J(Y). Unfortunately, the extraction of information is hampered by the
fact that prominent — and physically interesting — features in n(p) may
not be strongly reflected in J(Y) [6].
To see this, consider the two recent calculations of the ground state
n(p) discussed earlier and shown in Fig. 5(a). Both calculations predict
a condensate fraction, which appears as a delta function, with no=9.2
%. Both calculations also predict quite similar behavior at intermediate
and large p. However, they differ markedly at small p. The variational
calculation exhibits singular behavior due to coupling of long wave-
length density fluctuations (phonons) to the condensate which are not
present in the GFMC result, presumably due to finite size effects in the
calculation.
While n(p) for these two calculations is quite different, the correspond-
ing JIA(Y), shown in Fig. 5(b), is remarkably similar. The singular
behavior, which is the dominant feature in the variational n(p) at small
/?, is quite small in J(Y). When instrumental broadening is taken into
account the (now small) differences between the predicted scattering
for the two calculations all but disappear, as shown in Fig. 5(c). The
predictions of the two quite different n(p)s are now nearly indistinguish-
able ! In principle, the differences between the two different n(p)s are still
present in Fig. 5(c). In practice, a measurement of the scattering would
need fantastically good statistical accuracy to ever hope to observe these
differences.
This insensitivity to the details at small p is a direct consequence of
the IA. The Compton profile (9), which is proportional to the measured
scattering, is the momentum distribution in the direction of the momen-
tum transfer averaged over the longitudinal components. Since this is
the integral of pn(p) singular behavior at small p will be suppressed due
to the factor of p in the integrand. Alternately, features at large p will be
enhanced for exactly the same reason.
BEC in Liquid Helium 65
""1"":
0.5 0.5
— HNO/$ : : (c) A —HNC/$ :
dFMd
0.4 — / \ - 0.4
'— f
0.3 0.3
\ ~
'- \ -
0.2 — \ ~ h / \ H 0.2
"1"
0.1 - I \ - 0.1
0.0 0.0
-4 -2 4 - 4 - 2
Fig. 5. (a) Two theoretical calculations of n(p) in liquid helium at T = 0. The solid
curve is the GFMC result, and the dashed curve is the HNC/S variational result,
(b) The longitudinal momentum distributions described by J(Y) corresponding
to the two ground state n(p) in (a), (c) The combined effects of the convolution
of J(Y) in (b) with the instrumental resolution function and Silver's final-state
broadening.
0.1 —
0.0
-4
Y (A" 1 )
5 Experimental Results
The suggestion that neutron scattering measurements at large momentum
transfers could directly observe the condensate set in motion a number
of studies by a variety of investigators [37]. Unfortunately, none of
these studies has succeeded in reaching the original goal: a direct and
unambiguous observation of the condensate. In fact, with our current
understanding of these measurements and FSE which complicate them,
we now realize that a direct observation of a 3 -function component in
n(p) is unlikely in the foreseeable future [38]. However, these studies
of liquid helium have provided a wealth of detailed information on
the momentum distribution in this strongly interacting quantum liquid.
With the use of some well-founded theoretical results, information on
the magnitude of the condensate fraction can be extracted, including its
temperature and density dependence.
The earliest attempts to measure n(p) and no were carried out at reactor
based sources [39]. These measurements provided general information
on n(p) and pointed the way toward techniques to extract the condensate.
However, they were limited by the large FSE at the relatively low Q («
5-15 A" 1 ) attainable at reactor based sources. Some measurements were
carried out at larger Q, but only at the expense of resolution making it
difficult to obtain accurate information on n(p).
More recently, measurements [40, 41, 42, 43, 44] have been carried
out using spallation neutron sources, such as the Intense Pulsed Neutron
Source (IPNS) at Argonne National Laboratory. These sources have a
high flux of epithermal (high energy) neutrons allowing high resolution
measurements to be carried out at much larger Q than are obtainable
with reactors. The higher Q have the advantage that the scattering is
consistent with the predictions of the IA, in terms of the location and
symmetry of the observed scattering. In addition, FSE are more amenable
BEC in Liquid Helium 69
0.2 —
0.1 —
0.0
-4
Y (A" 1 )
Fig. 7. Observed scattering at T = 0.35 K. The line is a fit of the model scattering
function discussed in the text broadened by the instrumental resolution function
and the final-state broadening function of Silver [32].
0.00
-0.25
-0.50
-0.75
t
•- 4I , -, ,2, I , , , 0
3
, ! • , . , 21 , , . , 4I
1
Y (A- )
Fig. 8. Observed J(Y) at a variety of temperatures in the normal and superfluid
phases at a density of 0.147 g cm"3.
11
• • I" " I• • " I" "1
All
0.50
(A"1)
Fig. 9. JModei(^) with instrumental resolution and FSE deconvoluted obtained
from the observed scattering shown in Fig. 8.
to remove the effects of instrumental resolution and FSE and to invert the
transformation between n(p) and J(Y). However, as discussed previously,
these are ill-defined procedures [45] that are strongly affected by statistical
noise present in the experimental data. Therefore, rather than attempting
to deconvolute the instrumental resolution, it is more appropriate to fit
an expression for J(Y), broadened by instrumental resolution and FSE,
to the observed scattering data. Theoretical predictions for J(Y) are
discussed in the following section. Here we discuss a simple expression
for J(Y) which has sufficient flexibility to reflect the behavior of the
true scattering accurately, yet which is also sufficiently constrained so
that unphysical behavior is not introduced due to the finite statistical
accuracy of the data.
The scattering function that we have found most convenient [41] for
describing the observed scattering is a sum of two Gaussians
(12)
0.6 1 1
, , , , 1, , , i I
" I data "
T=0.35K
- (upper) 2d-t n 0 =10%-
0.4 —
: (a)
A - (middle) 2(3 +n0 =4%J
(lower) 2d+ n Q =0% -
0.2 -
]
0.0 ,1 i i , • \7 —
1
LTIWMI
1 1
-4 -2
- (upper) 2 d + n Q =0%
- (middle) 2 d + n Q = 4 % ,
- ( l o w e r ) 2 d + n Q =10% -
0.0
Fig. 10. (a) Three different two-Gaussian fits to the data at 0.35 K. (b) The
corresponding model momentum distributions JModei(Y). One is a two-Gaussian
fit with a condensate fraction of 10% , another is a two-Gaussian fit with
a condensate fraction of 4% , and the third is a two-Gaussian fit with no
condensate fraction.
6 Comparison to Theory
A direct model-independent determination of no is beyond current ex-
perimental capabilities for the reasons discussed in Section 5. However,
we may still obtain information on the condensate by comparing theo-
retical calculations of n(p) with the experimental data. Such comparisons
provide a direct test of the theoretical predictions and, indirectly, give
information about the magnitude of the condensate.
The dashed line in Fig. 11 (a) shows the theoretical prediction for JIA(Y)
in the normal liquid using the PIMC calculations [25] of n(p). The the-
oretical n(p) has been converted to J(Y) using the IA and broadened
by the instrumental resolution. The agreement between the theoretical
predication and the experiment is excellent. In this case the IA, calcu-
lated using the theoretical n(p)9 provides an excellent description of the
scattering in the normal liquid.
FSE have little effect on the observed scattering in the normal liquid
at these Q values. However, they may be included by convoluting the
theoretical predictions with the broadening function shown in Fig. 6. The
solid lines in Fig. 11 (a) and (b) are obtained when FSE are included.
Within the statistical accuracy of the measurements, there is no observable
change when FSE are included. Since, based on the co2 sum rule, FSE do
not change the second moment of the scattering, they have little effect
on the broad, nearly Gaussian J(Y) of the normal liquid.
74 P. Sokol
1 1 1 1 1
. " " I " I ' " I " : . ...1 . 0.5
0.5 - (a) I 3.5 K : I 0.35 K :
0.4 — jjjjfe
H
—
l — 0.4
J(Y) °-;
0.3
0.2
rF /A\ I
h /\
- / \
-
-
0.3
0.2
\
1 1 1 1 1
i i 1 i
0.1 0.1
0.0
0.0 ^fTMi.,,1 I I I -
-4 -2
Y (A"1)
Fig. 11. Comparison of the observed scattering in (a) the normal liquid at 3.5 K
and (b) the superfluid at 0.35 K, with the theoretical predictions from GFMC
and PIMC calculations. The dashed lines are the theoretical predictions with only
instrumental broadening included. The solid lines are the theoretical predictions
with instrumental resolution and final-state effects included.
The dashed line in Fig. ll(b) shows a similar comparison of the GFMC
calculations [21] to the scattering in the superfluid. The agreement
between the theoretical and experimental results is quite poor, particularly
in the region of the peak center, where the condensate would have the
largest contribution. Based on this comparison alone, which has neglected
FSE, we would conclude that no, if not identically zero, would have a
much smaller value than the theoretical predictions.
The inclusion of FSE in the comparison between the experimental and
theoretical results is essential in the superfluid phase. This is due to the
appearance of a sharp feature in n(p) for the superfluid: the condensate.
While FSE have little effect on the broad component of the scattering, as
observed in the normal liquid, they significantly broaden the contribution
from the condensate. Taking FSE into account, the agreement between
theory and experiment is now excellent! For the first time, ab initio
numerical calculations of n(p) in the superfluid are in good agreement
with experiment [40].
An important point regarding the final-state corrections is in order
here. The final-state broadening prediction of Silver, as shown in Fig. 6,
has a narrow central peak and negative tails at high Y. The negative tails
are essential if the broadening function is to satisfy the second-moment
BEC in Liquid Helium 75
sum rule. Thus, final-state effects not only broaden the condensate peak,
they also shift intensity throughout the entire spectrum. For the particular
form of the FSE used here, the negative tails will cause a depletion of the
scattering at intermediate Y when a condensate is present.
In view of the discussion in Section 5 regarding the relationship be-
tween J(Y) and n(p\ it is appropriate to examine how sensitive the
observed scattering is to the theoretical n(p). Inevitably there is a finite
statistical accuracy attached to the experimental results, and a whole
range of different n(p)s may give equally good agreement with the data.
If the statistical accuracy of the results is high then only a limited range
of n(p)s, all with very similar shapes, will be consistent with the data.
Alternatively, if the statistical accuracy is poor then the experimental
results will only place very weak constraints on the underlying shape of
n(p).
Unfortunately, it is not possible to vary the condensate fraction in
numerical calculations to see what effect it has on the comparison with
the experimental data. The numerical calculations use as input only the
known interatomic potential. Based on this potential, the full many body
state of the liquid is calculated. There are no parameters that one can
adjust to make no larger or smaller. The result of the simulation, just
as in the real system, is determined solely by the interaction between
the atoms. Short of adjusting the interaction, which would undoubtedly
change the entire n(p), we cannot adjust no.
We can, however, attempt to make a limited test of the sensitivity of
the experimental results to the value of the condensate fraction. The
momentum distribution can be decomposed into a condensate contri-
bution and a non-condensed contribution. We may adjust the relative
weights of the condensate component and the non-condensed compo-
nent such that they still satisfy the normalization of n(p). However,
once we choose a value for the condensate different from the theoret-
ical value the resultant momentum distribution no longer corresponds
to any ab initio calculation for helium. Thus, while we are testing the
sensitivity of the scattering to the value of no, it is only in the lim-
ited range where n(p) has a very special shape for the non-condensed
component.
With the above caveats in mind, we may replace the condensate d-
function with a Gaussian of variable width and amplitude, keeping the
shape of the non-condensate distribution the same as the theoretical
prediction. The best agreement is obtained when the width of the
Gaussian is less than 0.05 A" 1 and no is 10%, in agreement with the
76 P. Sokol
best numerical predictions for the condensate. Significant deviations are
observed when the width is greater than 0.2 A" 1 and no is less than 8%
or greater than 12%. For this particular model for the non-condensate
part of n(p) provided by the GFMC calculations, we find that there is
indeed a condensate with no — 10 ± 2%. However, we note that changing
the shape of the uncondensed n(p) would also have an effect on the value
for the condensate fraction.
In a similar fashion, the sensitivity to the expected singular behavior in
n(p) at small p can be examined. The GFMC results, which give excellent
agreement with the observed scattering, do not contain the expected
singular contribution. (Because of finite-size effects, these results are
only valid for p > TC/L, where L is the system size.) The variational
n(p), discussed previously, explicitly includes this behavior. Both results
are in excellant agreement with experiment. This is not very surprising,
since the weak singular behavior at small p is suppressed when n(p) is
transformed to J(Y), as discussed earlier. Thus, the predicted small p
singular behavior makes little contribution to the observed scattering
and, with the experimental techniques now available, will be difficult, if
not impossible, to observe.
Thus, the experimental results in the superfluid provide a clear indi-
cation of a narrow component in n(p) containing approximately 9-10%
of the intensity, which is precisely that expected for the condensate.
Unfortunately, due to the finite statistical error inherent in any exper-
iment, they cannot definitely prove the existence of a condensate in
the form of a <5-function. Some other singular behavior not associated
with a condensate could be responsible for the increase in the scatter-
ing at small p observed in the superfluid. As seen in the comparison
with the variational n(p), however, this would have to be very singular
behavior, much more than the 1/p singularity, to agree with the exper-
imental results. Thus, while the experimental results cannot rule out
a ground state n(p) which does not contain a S-function condensate,
they do provide strong evidence for a very narrow feature containing
10+2% of the total area. The excellent agreement with the numerical
results suggests that this very narrow feature is indeed due to the Bose
condensate [3].
Similar comparisons have been carried out at a variety of temperatures
and several different densities in both the normal and superfluid phases.
GFMC and variational calculations are used for comparison with low
temperature measurements and PIMC results are used for comparison
to measurements above 1 K. The agreement is excellent over the entire
BEC in Liquid Helium 77
temperature range! Theory and experiment appear to have converged
for n(p) in liquid 4 He at low densities (SVP).
15.0
12.5 -
10.0 -
2.5 —
0.0
10% is in excellent agreement with the GFMC result of 9.2%. The weak
temperature dependence of n0 seen in these fits at low temperatures is
also consistent with the PIMC results [25] and the finite-temperature
extensions of variational calculations.
The condensate wave function must vanish at the second-order
superfluid-normal fluid phase transition, since it is the microscopic order
parameter for the superfluid phase. According to the renormalization
group theory for second-order phase transitions, the temperature de-
pendence of rc0 near Tx is expected to take the form n0 = At2P, where
t = (1 - T/Tx). Unfortunately, the errors in the current experimental
estimates are too large to place any useful constraints on the critical
exponent which controls the behavior of the condensate close to the
superfluid transition. However, we may compare this prediction with our
experimental results using the accepted value of j8=0.35 obtained from
80 P. Sokol
12.5 -
D
2.5 -
nn
0.14 0.16
1 . . . , 1 , ,
0.18
I.,..
0.2
P (g/cc)
Fig. 13. The value of n0 as a function of density at 0.75 K obtained from fits
using the method of Snow et al. [50] (crosses). The GFMC results for n0 (squares)
and the HNC/S results (diamonds) are also shown.
8 Conclusions
References
[1] The cessation of bubbling as 4 He passed through the superfluid
transition was first noted by J. C. McLennan, H. D. Smith, and J.
O. Wilhelm, Phil. Mag. 14, 161 (1932).
BEC in Liquid Helium 83
[2] The first experiments showing superfluidity explicitly were reported
by J.F. Allen and A.D. Misener, Nature 141, 75 (1938), and P.
Kapitza, Nature 141, 74 (1938).
[3] F. London, Nature 141, 643 (1938).
[4] This method was first proposed by P.C. Hohenberg and P.M. Platz-
man, Phys. Rev. 152, 198 (1966).
[5] H.R. Glyde and E.C. Svensson, in Methods of Experimental Physics,
K. Skold and D. L. Price, eds. (Academic, N.Y., 1987), Vol.23, Part
B, p. 303.
[6] RE. Sokol, T. R. Sosnick, and W.M. Snow, in Momentum Distri-
butions, R. N. Silver and P. E. Sokol, eds. (Plenum, N.Y, 1989),
p.139.
[7] R.K. Pathria, Statistical Mechanics (Pergamon, Oxford, 1972).
[8] R.A. Aziz, V.P.S. Nain, J.S. Cerley, W.L. Taylor, and G.T. Mc-
Conville, J. Chem. Phys. 70, 4330 (1979).
[9] W.E. Keller, Helium-3 and Helium-4 (Plenum, N.Y., 1969).
[10] P. E. Sokol, R.N. Silver and J. W. Clark, in Momentum Distributions,
R.N. Silver and P. E. Sokol, eds. (Plenum, N.Y, 1989).
[11] el Hadi et al, Physica 41, 289-304 (1969); Phys. Chem. Ref. Data
2, 923 (1973); 6, 51 (1977).
[12] O. Penrose and L. Onsager, Phys. Rev. 104, 576 (1956).
[13] C.N. Yang, Rev. Mod. Phys. B 34, 694 (1962).
[14] P.W. Anderson, Rev. Mod. Phys. B 38, 298 (1966).
[15] G. Baym, private communication.
[16] J. W. Clark and M.L. Ristig, in Momentum Distributions, R. N. Silver
and P. E. Sokol, eds. (Plenum, N.Y, 1989), p. 39.
[17] A. Griffin, Can. J. Phys. 65, 1368 (1987).
[18] P.C. Hohenberg and P.C. Martin, Ann. Phys. (NY) 34, 291 (1965).
[19] J. Gavoret and P. Nozieres, Ann. Phys. (NY) 28, 349 (1964).
[20] E. Manousakis, V. R. Pandharipande, and Q. N. Usmani, Phys. Rev.
B 33, 7022 (1985); E. Manousakis and V. R. Pandharipande, Phys.
Rev. B 30, 5064 (1984); E. Manousakis, in Momentum Distributions,
R. N. Silver and P. E. Sokol, eds. (Plenum, N.Y, 1989), p. 81.
[21] P.A. Whitlock and R. Panoff, Can. J. Phys. 65, 1409 (1987); R.M.
Panoff and P.A. Whitlock, in Momentum Distributions, R. N. Silver
and P. E. Sokol, eds. (Plenum, N.Y, 1989), p. 59.
[22] S.A. Vitiello, K.J. Runge, G.V. Chester and M.H. Kalos, Phys. Rev.
B 42, 228 (1990).
84 P. Sokol
[23] E. Feenberg, Theory of Quantum Fluids (Academic, N.Y., 1969); C.E.
Campbell, Progress in Liquid Physics, C.A. Croxton, ed. (Wiley,
N.Y., 1968), ch. 6.
[24] C.E. Campbell, in Excitations in Two-Dimensional and Three-
Dimensional Quantum Fluids, A.G.F. Wyatt and H.J. Lauter, eds.
(Plenum, N.Y., 1991), p. 159.
[25] D.M.Ceperly and E.L.Pollock, Phys. Rev. Lett. 56, 351 (1986);
D.M. Ceperly, and E.L. Pollock, Can. J. Phys. 65, 1416 (1987);
E.L. Pollock and D.M. Ceperley, Phys. Rev. B 36, 8343 (1987);
D.M. Ceperley, in Momentum Distributions, R.N. Silver and
RE. Sokol, eds. (Plenum, N.Y, 1989), p. 71.
[26] A.D.B. Woods and V.F. Sears, Phys. Rev. Lett. 39, 415 (1977); H.A.
Mook, Phys. Rev. Lett. 32, 1167 (1974).
[27] D.L. Price and K. Skold in Methods in Experimental Physics, Vol.
23A, D.L. Price and K. Skold, eds. (Academic, N.Y, 1987), p. 1.
[28] RE. Sokol, Can. J. Phys. 65, 1393 (1987).
[29] J.J. Weinstein and J.W. Negele, Phys. Rev. Lett. 49, 1016 (1982).
[30] R.N. Silver, in Momentum Distributions, R.N. Silver and RE. Sokol,
eds. (Plenum, N.Y, 1989), p. 111.
[31] R. Feltgen, H. Kirst, K.A. Koehler, and F. Torello, J. Chem. Phys.
26, 2630 (1982).
[32] R.N. Silver, Phys. Rev. B 37, 3794 (1988); Phys. Rev. B 39, 4022
(1989).
[33] H. A. Gersch and L. J. Rodriguez, Phys. Rev. A 8, 905 (1973); L.
J. Rodriguez, H. A. Gersch and H. A. Mook, Phys. Rev. A 9, 2085
(1974).
[34] C. Carraro and S.E. Koonin, Phys. Rev. Lett. 65, 2792 (1990); Phys.
Rev. B 41, 6741 (1990).
[35] P.E. Sokol, to be published.
[36] P.E. Sokol, T.R. Sosnick, W.M. Snow, and R.N. Silver, Phys. Rev.
B 43, 216 (1991); K.W. Herwig, P.E. Sokol, W.M. Snow, and R.C.
Blasdell, Phys. Rev. B 44, 308 (1991).
[37] For a brief review, see E.C. Svensson, in 75th Jubilee Conference on
Helium-4, J.G.M. Armitage, ed. (World Scientific, Singapore, 1983),
p. 10.
[38] P.E. Sokol and W.M. Snow, in Excitations in Two-Dimensional and
Three-Dimensional Quantum Fluids, A.G.F. Wyatt and H.J. Lauter,
eds. (Plenum, N.Y, 1991) p. 49.
[39] R.A. Cowley and A.D.B. Woods, Phys. Rev. Lett. 21, 787 (1968).
BEC in Liquid Helium 85
[40] T.R. Sosnick, W.M. Snow, RE. Sokol, and R.N. Silver, Europhys.
Lett. 9, 707 (1989).
[41] T.R. Sosnick, W.M. Snow, and RE. Sokol, Phys. Rev. B41, 11185
(1990).
[42] K.W. Herwig, RE. Sokol, T.R. Sosnick, W.M. Snow, and R. C.
Blasdell, Phys. Rev. B 41, 103 (1990).
[43] K.H. Anderson, W.G. Stirling, A.D. Taylor, S.M. Bennington, Z.A.
Bowden, I. Bailey, and H.R. Glyde, Physica B 180, 865 (1993).
[44] J. Mayers and A.C. Evans, Rutherford Appleton Laboratory Report
RAL-91-048.
[45] D.S. Sivia and R.N. Silver, in Momentum Distributions, R.N. Silver
and RE. Sokol, eds. (Plenum, N.Y., 1989), p. 377.
[46] H. W. Jackson, Phys. Rev. A 10, 278 (1974).
[47] V.F. Sears, E.C. Svensson, P. Martel and A.D.B. Woods, Phys. Rev.
Lett. 49, 279 (1982).
[48] A. Griffin, Phys. Rev. B 32, 3289 (1985).
[49] J.A. Lipa and TCP. Chui, Phys. Rev. Lett. 51, 2291 (1983).
[50] W. M. Snow, Y. Wang and P. E. Sokol, Europhys. Lett. 19, 403
(1992).
[51] D.S. Greywall and G. Ahlers, Phys. Rev. A 7, 2145 (1973).
[52] V.F. Sears, Can. J. Phys. 63, 68 (1985).
[53] G. J. Hyland, G. Rowlands, and F. W. Cummings, Phys. Lett 31A,
465 (1970); F. W. Cummings, G. J. Hyland, and G. Rowlands, Phys.
Lett 86A, 370 (1981).
[54] G.L. Masserini, L. Reatto, and S.A. Vitiello, J. Low Temp. Phys. 89,
335 (1992).
[55] H.R. Glyde and W.G. Stirling, in Momentum Distributions, R.N. Sil-
ver and RE. Sokol, eds. (Plenum, N.Y, 1989), p. 123.
Sum Rules and Bose-Einstein
Condensation
S Stringari
Dipartimento di Fisica
Universitd di Trento
1-38050 Povo
Italy
Abstract
Various sum rules, accounting for the coupling between density and particle
excitations and emphasizing in an explicit way the role of Bose-Einstein
condensation, are discussed. Important consequences on the fluctuations of
the particle operator as well as on the structure of elementary excitations
are reviewed. These include a recent generalization of the Hohenberg-
Mermin-Wagner theorem holding at zero temperature.
1 Introduction
The sum rule approach has been employed extensively in the literature
in order to explore various dynamic features of quantum many body
systems from a microscopic point of view (see [1] and references therein).
An important merit of the method is its explicit emphasis on the role of
conservation laws and of the symmetries of the problem. Furthermore,
the explicit determination of sum rules is relatively easy and often requires
only a limited knowledge of the system. Usually the sum rule approach
is, however, employed without giving special emphasis to the possible
occurrence of (spontaneously) broken symmetries. For example, the
most famous /-sum rule [2] holding for a large class of systems is not
affected by the existence of an order parameter in the system.
The purpose of this paper is to discuss a different class of sum rules
which are directly affected by the presence of a broken symmetry. These
sum rules can be used to predict significant properties of the system which
are the consequence of the existence of an order parameter. In this work
we will make explicit reference to Bose systems and to the consequences
of Bose-Einstein condensation (BEC). Most of these results can however
86
Sum Rules and Bose-Einstein Condensation 87
be generalized to discuss other systems exhibiting spontaneously broken
symmetries.
In Section 2 we use a sum rule due to Bogoliubov in order to derive
important constraints on the fluctuations of the particle operator as
well as to obtain Goldstone-type bounds for the energy of elementary
excitations. In Section 3 we explore the consequences of BEC on the long
range behavior of the half diagonal two-body density matrix and discuss
the coupling between the density and particle pictures of elementary
excitations in Bose superfluids.
/:
is the most remarkable among the sum rules depending explicitly on the
order parameter n0 = jj | (a0) |2. In (1) s^A^B{o}) is the spectral function
= / )-d(o (5)
J—o 0)
and the result ({A^9A}) > 2UBTXA^A following from the fluctuation
dissipation theorem.
Choosing in the Bose case A* = aq and B = p q we obtain the inequal-
ity [5]
that can be easily calculated at small q. In fact, in this limit, the inverse
energy weighted sum rule (compressibility sum rule) / ^S(q,a>) = ^x(#)
approaches the compressibility parameter of the system (limq-+ox(cl) =
1/mc2), while the energy weighted sum rule (/-sum rule) is given by
(7). As a consequence S(q) < q/2mc at small q and the momentum
distribution n(q) diverges at least as
n(q) = nor^. (11)
It is worth pointing out that the l/q law for the momentum distribution,
already known in the literature [9, 10], has been obtained here without
any assumption on the nature of the elementary excitations of the system
(phonons in a neutral Bose superfluid) and follows uniquely from the
existence of BEC, the validity of the /-sum rule and the finiteness
of the compressibility. By imposing the proper normalization on the
momentum distribution one can then rule out the existence of BEC in
one-dimensional systems at zero temperature. Note that the ideal Bose
Gas does not violate the theorem since in this case the compressibility
sum rule for x(q) diverges as l/q2.
Starting from the same inequality (8), it is also possible to rule out the
existence of long range order in other one-dimensional systems at T = 0
90 S. Stringari
(isotropic antiferromagnets, crystals) [8]. The theorem does not apply
to one-dimensional ferromagnets since in this case the inverse energy
weighted sum rule (magnetic susceptibility) diverges as 1/q2.
Inequality (9) has been used recently [11] to rule out the existence of
BEC in the bosonic representation of the electronic wave function
^ . , ^ ) (12)
Uj
2 ([A\[H,A]])([B\[H,B]])
M
°- \(\A\B\)\2 ' ( }
Bounds of the form (14) were first considered by Wagner [4]. Result (14)
can be obtained using the inequality (holding at zero temperature)
where the rhs coincides with the ratio between the energy weighted and
the inverse energy weighted sum rules relative to the operator A. Use of
the Bogoliubov inequality (4) then yields (14).
Result (14) has the important merit of providing a rigorous upper
bound for a>o in terms of quantities involving commutators. This is an
advantage, for example, with respect to the so-called Feynman bound:
w 2(
° ^-;ri^ [ i^ + N F W p
(17)
where we have carried out the commutators using the grand canonical
hamiltonian H' = H — fiN (ja is the chemical potential) and taken a
central potential with Fourier transform V(p). The quantitities n(p) and
n(p) are defined by n(p) = (ajflp) and n(p) = \{a\aLv +a p a_ p ).
Result (17) is a rigorous inequality holding for any Bose system inter-
acting with central potentials. It has the form of a Goldstone theorem.
In fact, since its rhs behaves as q2 when <gr —>• 0, it proves the existence
of gapless excitations, provided the order parameter no is different from
zero.
It is worth noting that in the limit of a dilute Bose gas (/i = NV(Q\
92 S. Stringari
n(p) = n(p) = iV<5p,o,no = 1)> this upper bound coincides with the Bogoli-
ubov dispersion law
~2 r »i
(18)
for any value of q. This result follows from the fact that the sum rules
entering inequality (14) are exhausted by a single excitation, multi-particle
states playing a negligeable role in a dilute gas. Of course in a strongly
interacting system, such as liquid 4 He, multiparticle excitations are much
more important and the bound (17) turns out to be significantly higher
than the lowest excitation energy coo(q).
It is useful to compare the Goldstone-type inequality (17) with another,
also rigorous, upper bound, which is derivable from (14) by choosing
i2 ni. int. n
m m
where (EK) is the ground state kinetic energy and g(s) is the pair corre-
lation function. The bound (20) then becomes:
on the sum rules discussed above. A first important result is that the
Goldstone-type upper bound (17) is not directly affected by the external
field because of the exact commutation property
(25)
Vice versa, the cubic energy weighted sum rule for the density operator
pq obtains an extra contribution from the external force given by
(26)
At small q the new term provides the leading contribution to the triple
commutator (21). Since the quantity ([p_q, [H,pq]]) is not changed by
the external force, being still given by the /-sum rule (7), the upper
bound (20) no longer vanishes with q. This different behaviour is due
to the fact that the current jq=o is not conserved in the presence of
the external field (24). This result also reveals that the relationship
jq = qy/Nno/2m(aq — a_q) between the current and the gradient of the
phase operator, holding in a translationally invariant dilute Bose gas, is
no longer valid in the presence of an external potential and gives rise
to a normal (non-superfluid) component of the density of the system at
zero temperature. This is discussed, for example, in Huang's article in
this volume.
94 S. Stringari
3 The Half Diagonal Two-body Density Matrix
In this section we discuss another sum rule, also sensitive to the presence
of Bose-Einstein condensation, given by [18]
p +co i
a
/ 1 _ -B -q W> (27)
J—
where si is the spectral function already introduced in Section 1 with
the choice A* = aq + a!q and £ = pq.
In contrast to (1), which describes the commutation relation between
the density and the phase operators, this sum rule cannot be expressed in
terms of a commutator. Physically it accounts for the coupling between
the density of the system and the modulus of the condensate, proportional
to aq + ai q . An interesting feature of this sum rule is that it fixes the long
range behaviour of the half diagonal two-body density matrix
P(2)(ri,r2;ri,r2) = ( ^ ( r O ^ m ^ m ^ ) ) . (28)
The occurrence of BEC is in fact not only relevant for the long-range
behaviour of the one-body density matrix
(29)
fixed by the condition
\imp(1\r,r') = pn0, (30)
r'->oo
but also for the one of the two-body matrix (28). The long-range order
(LRO) in the two-body density matrix (28) is naturally defined by [19]
lim p (2) (ri,r 2 ;ri,r 2 ) = n0p2(l + Fx(\ n - r 2 |)) (31)
= pJ drV(r)(l+F (r))
1 (36)
relating the LRO function F\(r) to the chemical potential of the system.
The result in (36) holds for any system exhibiting Bose condensation and
interacting with central potentials.
The LRO function F\(q) plays a crucial role in the coupling between
the density
and particle
\P) - ^ - , 1 0 ) (38,
(39)
The coupling turns out to be complete when q -• 0 (in fact in this limit
one has S(q) = q/2mc,n(q) = nomc/2q and F\(q) = —^), showing in
96 S. Stringari
an explicit way that in the hydrodynamic limit the density and particle
picture of elementary excitations of a Bose condensed system coincide.
The coupling between the density and particle pictures is at the basis
of fundamental properties exhibited by Bose superfluids (for a recent
exhaustive discussion see Ref. [21]).
It is finally interesting to compare the average excitations energies of
the states (37) and (38). Both energies provide a rigorous upper bound
to the energy of the lowest excited state of the system. The energy of the
Feynman state (37) is given by the famous result [20]
_p = ^4 - — - - ( 4 3 )
€P=ii-
—((E K) + 2{V))9 (44)
N
where (EK) and (V) are the kinetic energy and the potential energy
Sum Rules and Bose-Einstein Condensation 97
relative to the ground state. At zero pressure, where \i = j^((EK) + (V)),
(44) yields eP = -(V)/N = 21 ~ 22K in superfluid 4 He. It is instructive
to compare the above value with the corresponding average of the
Feynman energy:
(45)
References
[I] S. Stringari, Phys. Rev. B 46, 2974 (1992).
[2] D. Pines and Ph. Nozieres, The Theory of Quantum Liquids (Ben-
jamin, New York, 1966), Vol.1; Ph. Nozieres and D. Pines, The
Theory of Quantum Liquids (Addison-Wesley, 1990), Vol.11.
[3] N.N. Bogoliubov, Phys. Abh. SU 6, 1 (1962).
[4] H. Wagner, Z. Physik 195, 273 (1966).
[5] PC. Hohenberg, Phys. Rev. 158, 383 (1967).
[6] N.D. Mermin and H. Wagner, Phys. Rev. Lett. 17, 1133 (1966).
[7] N.D. Mermin, Phys. Rev. 176, 250 (1968).
[8] L. Pitaevskii and S. Stringari, J. Low Temp. Phys. 85, 377 (1991).
[9] T. Gavoret and Ph. Nozieres, Ann. Phys. (NY) 28, 349 (1964).
[10] L. Reatto and G.V. Chester, Phys. Rev. 155, 88 (1967).
[II] L. Pitaevskii and S. Stringari, Phys. Rev. B 47, 10915, (1993).
[12] S. Girvin and A. MacDonald, Phys. Rev. Lett. 58, 1252 (1987).
[13] R.B. Laughlin, Phys.Rev.Lett. 50, 1395 (1983).
[14] S.C. Zhang, Int. J. Mod. Phys. 6, 25 (1992).
[15] W. Kohn, Phys.Rev. 123, 1242 (1961).
[16] R.D. Puff, Phys. Rev. A 137, 406 (1965).
[17] D. Pines and C.-W. Woo, Phys. Rev. Lett. 24, 1044 (1970).
[18] S. Stringari, J. Low Temp. Phys. 84, 279 (1991).
98 S. Stringari
[19] M.L. Ristig and J. Clark, Phys. Rev. B 40, 4355 (1989).
[20] A. Bijl, Physica 8, 655 (1940); R.P. Feynman, Phys. Rev. 94, 262
(1954).
[21] A. Griffin, Excitations in a Bose-Condensed Liquid (Cambridge Uni-
versity Press, 1993).
Dilute-Degenerate Gases
F. Laloe
Laboratoire de Physique de I'ENS'f
24, rue Lhomond
F75231 Paris, Cedex 05
France
Abstract
99
100 F. Laloe
without requiring any new ingredient, such as some variational wave func-
tion adequate for each channel of condensation.
1 Introduction
Recent progress in experiments looking for Bose-Einstein condensation
(BEC) in dilute systems, in spin polarized hydrogen [1, 2, 3], laser-cooled
clouds of atoms [4] and excitons[5, 6, 7], provides a natural stimulus
towards a more detailed understanding of the properties of degenerate
quantum gases. In these gases at very low temperatures, unusually large
values of the thermal wavelength XT may be obtained, and one can
reach situations where nX\ is of order one while the condition na3 <C 1
is still satisfied (n is the number density of particles and a is the range
of the interatomic potential); these systems are degenerate, but they
remain dilute in terms of the interactions. A motivation for this work is
the hope that BEC of these "dilute-degenerate gases" will be observed
experimentally and studied in detail in the near future.
A second motivation arises from the fact that dilute systems, such
as those studied in atomic physics, often provide useful model systems,
allowing calculations that give general views on physical mechanisms and
processes can be extended to denser systems. But the spirit of atomic
physics also implies a detailed treatment of two-body correlations (as
opposed to mean-field approximations); it is well-known that binary
collisions can be treated exactly, and there is no special reason to believe
that the occurrence of degeneracy will destroy this possibility. Moreover,
in the study of the Bose-Einstein and superfluid transitions, or processes
occurring in the build up of degeneracy, a microscopic approach may
be particularly useful because, after all, several basic questions in this
domain are still not answered very clearly. For instance, even at the
level of the one-body density operator, one can ask precisely how the
momentum distribution function is affected when one adds repulsive
hard cores to an ideal Bose condensed gas. By a continuity argument,
it is clear that the distribution will resemble that of the ideal gas, which
contains a delta function of momentum centred at zero velocity. But
two different scenarios are then conceivable: either the delta function
will survive the effects of the interaction, just changing its weight, or it
will acquire a finite width and become a "non-monochromatic peak".
The first scenario is commonly accepted, and indeed at zero temperature
there is a general argument by Penrose and Onsager [8] that proves its
validity; the argument basically uses only the fact that the ground state
Dilute-Degenerate Gases 101
wave function of a system of bosons has no node and, moreover, gives a
good order of magnitude estimate of the "condensate fraction" in liquid
4
He. Nevertheless, at non-zero temperature, the argument does not apply
anymore since the N-body wave functions of the excited states oscillate
around zero, and the existence of a momentum delta function becomes
less clear. Generally speaking, in most treatments of the Bose-Einstein
condensation in dilute systems, the zero width of the peak is not proved,
but, rather, it is assumed. This is in particular true of all first order
perturbation calculations, or in the various forms of mean-field theories
where, by construction, the wave functions remain unaffected under the
effect of the interactions, and certainly do not vanish on any of the
contact hypersurfaces mentioned in footnote.f The zero width of the
peak at the origin is more an assumption than a result.^
If one now goes to the level of two-particle density operators (correla-
tions), more questions arise, maybe more important conceptually because
they are related to the exact relation between the Bose-Einstein conden-
sation and superfluidity. We know that the former occurs in ideal gases
while the latter does not, but in interacting systems are they fundamen-
tally the same phenomenon, always taking place at exactly the same
temperature? Or is it conceivable that, for some class of systems, two
distinct transitions will occur? The very notion of superfluidity implies
that there is a mobile condensate that is robust against the effects of
dissipation, and one would like to understand in detail what kind of cor-
relations, occurring presumably at the same time in the space of momenta
and coordinates of the particles, physically create this robustness.
Of course these questions have been discussed in the literature, and
there is already an impressive body of work on the properties of the
Bose-Einstein and superfluid transitions in dilute systems. This includes
various contributions made after the important ideas introduced by Bo-
goliubov [10], the method of the pseudopotentials developed by Huang,
Lee, Yang and collaborators [11, 12, 13], the "binary collision approxima-
tion" of Lee and Yang [14, 15], just to refer to some of the best-known
t Actually, reasoning in the configuration space of the system, where hard cores make all
iV-body wave functions of the system vanish on a large number of contact hypersur-
faces [11], gives the impression that the peak centred at zero velocity might indeed have
a finite width: one could expect some kind of Heisenberg relation between the distance
over which the wave function can propagate in a 3N dimension space without vanishing
(of the order of the classical mean-free-path for hard cores in the system) and the width
of the momentum peak at the origin. If that were the case, the peak would not be really
infinitely narrow, except in the limit of vanishing densities.
J For a critical discussion of the experimental evidence of a condensate fraction in
superfluid helium 4, see [9].
102 F. Laloe
contributions. In most cases the present work will, of course, not do
more than recover the same physical results from another point of view,
where two-body correlations are emphasized in more detail. In par-
ticular, in our effort to treat short range correlations as precisely as
possible, we will avoid any perturbation series in the interatomic po-
tential V, even if resummations are possible, for instance by selecting
ladder diagrams in order to reconstruct the T collision matrix from V.
Rather, we shall use cluster expansions which, as is well-known (see,
for instance, [16]), are a tool of choice for obtaining density expansions
instead of V expansions; they contain exponential expressions of the
hamiltonians which do not create any divergence problem for infinite
potentials (hard cores). For this purpose, we will make use of generalized
quantum Ursell functions, actually Ursell operators, which intrinsically
do not contain statistics (as opposed to the usual Ursell functions, which
are already symmetrized [16, 17]): their action is defined, not only in the
space of states of the real system of indistinguishable particles, but also
in the larger space of an auxiliary system of distinguishable particles. If
the system is dilute in terms of the interactions but not of statistics, it
is possible to limit the calculation to low order Ursell operators! while,
with the usual (fully symmetrized) Ursell functions, the order would tend
to diverge with increasing degeneracy.
This method means that we have to give up the formalism of sec-
ond quantization, despite its well-known usefulness, but this is the price
to pay for the introduction of interactions and statistics in completely
independent steps. It also implies that, at some point, an explicit sym-
metrization of the wave functions becomes indispensable and, moreover,
that no approximation whatsoever can be made at this step: otherwise
the possibility of treating degenerate systems would be lost. Fortunately
it turns out that the symmetrization operation can indeed be performed
exactly without great difficulties; this is done by introducing simple prod-
ucts of operators, corresponding to exchange cycles, or, more generally,
simple functions of operators that correspond to summations over the
size of cycles up to infinity. In this way, interactions and degeneracy are
treated at completely different orders - actually all orders for the latter
- and relatively simple and concise expressions of the partition function
are obtained. When applied to dilute systems (with respect to the inter-
actions), the method can be seen as complementary to the method of
t For instance, when a Bose-Einstein condensation takes place, the Ursell operators in
themselves have no singular variation.
Dilute-Degenerate Gases 103
Belyaev for bosons [18] and that of Galitskii for fermions [19], extended
to non-zero temperatures.
The study of the Bose-Einstein condensation process itself in its various
forms (condensation of particles, of pairs, etc.) is delicate, because the
phase transition introduces a sensitivity to terms that would otherwise
remain small corrections. We propose to include them within a self
consistent equation for the "dressed Ursell operator". This leads to
results that remain mathematically simple and physically satisfying; for
instance, the repulsive part of the interactions tends to increase the value
of the chemical potential at which the condensation takes place; more
generally one determines qualitative predictions on the effects of the
repulsive or attractive part of the potential on the critical density. The
modification of the Ursell operator is somewhat similar to a Hartree-
Fock theory, but it is free of divergences for hard core potentials since it
incorporates short range correlations; it also includes some of the physics
contained in the non-linear Schrodinger equation [20], for instance the
modification of the condensate wave function under the effects of the
interactions. Finally, we will also see how condensation of pairs, triplets,
etc., appears naturally in this point of view. The present text does not give
the details of the calculations, which will be published elsewhere [21]; it
just attempts to summarize the physical ideas and the main results.
pi oc pF + r\pF. (5)
The reason is that the two terms that are linear in the exchange operator
turn out to take the simple form of squares of pp. If, for simplicity, we
assume that the particles do not have internal states, pF can be defined
through its Wigner transform pj?(r,p), and it easy to show that the
Wigner transform of relation (4) is
4 J.^-»') (8)
The former has no limitation concerning the distance at which the two
particles can be found, while the integral over k in the latter puts a
limit on this distance (of the order of the Planck constant divided by
the spread over relative momenta of the two particles); in other words
the exchange term has a range for the relative position which is of the
order of the De Broglie wavelength. This shows that the interpretation
of pr is simple: while pj gives the "full" one particle density operator,
defined independently of the position of the second particle, the operator
pr characterizes a "free" particle when it is far from the other (too far
away for exchange effects to play a role).
The calculations which lead to equation (5) also apply when the
particles have any kind of internal state. For instance one can assume
that the particles have (21 + 1) spin states, and that they are unpolarized
(equal probability to be in any of these states). It is then easy to see that
the exchange term is reduced by a factor (2/ + 1), which corresponds to the
reduction of effects of exchange when internal states of the particles can
be used to "tag" them. As expected, the factor disappears if all particles
are in the same spin state: all particles are then fully indistinguishable.
P/(l) = [ZFN]~\ Tr ( S
pF(l) <g> pF(2) <g>... 0 pF(N)S ) (9)
2,3,...,iV ^ J± si J
Y^kmk=N. (13)
k
The trace over the N particles in (9) is taken in a space that is simply the
tensor product of N single particle spaces of state, so that every cycle of
lengthfcin (12) corresponds to a trace taken in a different subspace which
introduces a separate factor FV Because the numbering of the particles
does not affect the value of F^ (it just changes the names of dummy
variables), we can for convenience renumber the relevant particles from
1 tofc.The effect of Q is then to move particle 1 into the place initially
occupied by particle 2, particle 2 into the place occupied by particle 3,
and so on, until one comes back to the place of particle 1. Introducing
a complete set of states {| 1 : cpn >} in the space of states^ of particle 1,
t In (10), one can suppress one of the operators S (or A) inside the braces by using
a circular permutation of the operators under the trace, combined with the relation
S2 = S. In (9) the trace is only a partial trace so that the argument does not apply
anymore; but, because all one-particle operators inside the brace are the same, their
product commutes with 5 (or A), so that the same simplification can be made.
J If the particles have internal states, the index n symbolizes at the same time the orbital
quantum numbers as well as those characterizing the internal state. For instance, if the
particles have spin /, a summation such as J^ n contains in fact two summations, one
over orbital quantum numbers, and a second over (2/ + 1) spin states.
Dilute-Degenerate Gases 107
* {mk} k
This formula uses the fact that cycles of an even number of particles
have parity —1 so that the parity of a cycle of length k is (— l)l+k. For
instance, if m\ = N (all the other nik being equal to zero), there is only
one permutation (the identity) that fits into this decomposition, and c is
equal to one. The other extreme happens when mN = 1 (all the other
m/c being equal to zero): one is then dealing with all cycles of maximum
length N, and it is easy to see that their number is (N— 1). More generally,
if we choose any combination of m^ obeying (13), we can distribute in
AM different ways the N particles into the N places contained inside
cycles, but all distributions do not generate different permutations P a for
two reasons: first every group of cycles of length k can be found in m&!
different orders; second, inside each cycle, any of the k particles can be
put arbitrarily in the first position, generating k different ways to recover
the same permutation. Therefore:
or:f
j e LogZj^ = iVTr{| (p)((p | p,} = N(q> \ p, | q>) (22)
(all derivatives with respect to e are taken at the value e = 0). If the grand
canonical ensemble is used, the N particle density operator is defined by:
If we insert (21) into this equation and compare with (23), we see that
this sum over N is nothing but the product of Z^c by the sum of the
contributions of every subspace N = 1,2,3,.... of the Fock space to the
t Here we take the convention where the operator pj is normalized to the number N of
particles (instead of one).
Dilute-Degenerate Gases 109
average value of a "one-particle operator", as usually defined in second
quantization. But this average value can also be expressed as a function
of the one particle density operator in the grand canonical ensemble pf
(normalized to the average number of particles), and we obtain:
1 - Y\Zpp
The two particle density operator may be obtained by a similar method
from a variation of pp of the form:
(29)
where S applies for bosons while A applies for fermions, and where the
operators KM are denned by:
Xi = exp-jBHo(l)
K2 =
KN =
in an obvious notation. We use cluster techniques to expand the operators
Kt into Ursell operators \J\ according to:
Kt = l/i
K2(l,2) =
K 3 (l,2,3) =
+C/2(3,l)C/i(2) +173(1,2,3).
(31)
In terms of the Ursell operators, the N particle operator KN can be
written in the form:
,.)... (32)
m D
\ i) \ / m\ factors m'2 factors
~~ ni (33)
Dilute-Degenerate Gases 111
The second summation corresponds to all non-equivalent ways to dis-
tribute the N particles into the variables of the Ursell operators, symbol-
ized by dots in (32). The only difference between our definitions and those
of Section 4.2 of [16] is that we use operators instead of symmetrized
functions; these operators are defined, not only within the state space
that is appropriate for bosons or fermions, but also in the larger space
obtained by the tensor product that occurs for distinguishable particles.
Hence the need for an explicit inclusion of S or A in (29).
As in the preceding section we decompose these operators into a sum
of permutations P a that, in turn, we decompose into independent cycles
of particles C:
^ ^ ] - . (36)
' {Pa} {U} diag
The next step is to remark that identical diagrams appear not only in
the same term of the double summation but also in many different terms.
Therefore, if Yl{mdia} symbolizes a summation over all possible ways to
decompose N according to (37), we can also write:
Z C
" = Jfi E ( " W >< f t [r*agnag ' (38)
' {wdiag} diag
Then a great simplification occurs because the sums over mdiag» which
are now independent, contain factorials according to (39) which intro-
duce simple exponentials; moreover the factors Q^N can be included by
multiplying every number Fdiag by e^"diae, so that:
(43)
r 2>1 = Tr
(45)
fTr
or:
r2>i = < 1 : q>Hl \< 2 : q>ni | l/ 2 (l,2) | 1 : <pni > | 2 : (pn2 >
<pni
(46)
that is:
Tr (47)
Similarly, one would calculate a contribution I\2 arising from the ex-
change of particles 2 and 3, and obtained by replacing U\(l) by U\(2) in
(47). More generally, when V2 clusters together k\ particles, belonging
to the same permutation cycle of length k\, with /C2 particles belonging
to another cycle of length fc2, the calculation of the effect of each of
these cycles remains very similar to that of Section 2.2: now we have
two particles that exchange separately with others, but the algebra of
Dilute-Degenerate Gases 115
operators remains the same for each of them. We therefore obtain the
contribution:
Tr (48)
For this class of diagrams, the value of the counting factor 1// is simply
1/2; this is because, while the numbering of the particles inside (72
can be used to distinguish the various clusters corresponding to the
same diagram, (72(m,n) is not distinct from l/2(n, m). Now the final step,
according to (42), is to make a summation over all possible values of k\
and fc2 after inserting an exponential of /? times the chemical potential
multiplied by the number of particles contained in the diagram. We
denote the sum by:
(49)
For the second class of diagrams, which we will call exchange diagrams,
the two particles contained in (72 are intermixed inside the same circular
permutation, which slightly complicates the situation. The first exchange
diagram corresponds to the two particles in question contained in the
same transposition:
^f = Tr{L/ 2(l,2)C 2(l,2)}=Tr{L/ 2(l,2)Pex}. (51)
J {2(,)e^i()} (55)
We will not give further details here, but from the preceding equations
it is not difficult to see that the generic term of this second class of
diagrams will be obtained from (48) by a simple replacement of U2 by
the product UjPex- Finally, the value of the grand potential (multiplied
by — /?), to first order in U2, will be
with
t We choose the origin of the energies so that the energy of the ground state is equal to
zero.
118 F. Laloe
the size L of the box containing the gas becomes very large.f When L
increases, the crossover becomes sharper and sharper and, in the ther-
modynamic limit, a singularity occurs at \i = 0, which is the origin of the
Bose-Einstein transition. Clearly, its very existence is related to the slow
divergence of a logarithmic function at the origin.
Before Bose-Einstein condensation takes place, the second term in the
right hand side of (57) remains an arbitrarily small correction if the range
of the interaction potential is reduced. Nevertheless, however small the
new term is, it completely changes the nature of the transition. From the
beginning, it is obvious that it does not contain logarithms, but fractions
which diverge more rapidly at the origin. It remains finite when \i —• 0,
because the behaviour at the origin of the integral involved in the trace
is of the type:
(i) The second particle in the added U2 may belong to a new, third
exchange cycle of length ki; this will correspond to a "direct-direct
diagram". To calculate its contribution, one can use reasoning that
is similar to that of Section 3.2, basically because the two U2 func-
tions play a symmetrical role in the diagram. One therefore expects
contributions where, in (48), U\ will be replaced as follows:!
f All the present discussion is more intuitive than rigorous; in particular the presence of
additional factors 1/2 in (61) and (62), arising from the counting factors / of the new
diagrams, is not obvious and requires a detailed study [21].
120 F. Laloe
(iii) The new Ui may also reconnect the two initial cycles, creating a "C/2
direct loop" and generating a term that can be expressed as a trace
over two particles only (and not three as in the two preceding cases).
For the moment we leave aside this class of diagrams.
Now, if we start from the exchange diagrams, the situation is similar,
and there are also three possibilities:
(i) Adding a new cycle again gives a "direct-exchange" diagram.
(ii) One can also lengthen the initial diagram and, as in (ii) above, introduce
a new U2 that contains two particles; if these particles are not entangled
with the two particles in the other U2 (running along the permutation
cycle, one obtains the two particles of one of the Ui functions and
then the two of the other), we obtain what we will call an "exchange-
exchange diagram".
(iii) Finally, one can do as in (ii) but entangle the two couples of particles
inside the JJi functions (along the cycle one gets alternatively particles
belonging to each 1/2; this provides a diagram with a "I/2 exchange
loop" which we will also leave aside for the moment.
We therefore see that substitutions (61) and (62) again apply. If now
we group these results and sum over all values of £3, we obtain:
, 2 ) [l+Pex] 1 ^ }
It is useful to notice that, in this equation, the "dressed operator" JJ\ has
an implicit definition, since it also appears in the denominator of (64);
in other words, U\ is defined by a self-consistent equation.
Dilute-Degenerate Gases 121
Now, when this substitution is done, the effects of the interactions
are contained in U\ itself and the value of the grand potential can be
obtained by the same calculation as for an ideal gas, but just replacing
U\ by U\. This gives the result|, similar to (43):
{[2(,)]} (67)
or, more generally, with two cycles of equal length k where all particles
are linked two by two by U2 functions:
{[2(,)]} (68)
Dilute-Degenerate Gases 123
None of the above terms contains an n factor, even for fermions, because
two cycles of the same parity are always involved.
Now, starting from the diagram that we called the "U2 exchange loop"
in the preceding section, we can proceed in a similar way as before, but
this time we do not consider two cycles of equal length, but only one
cycle of even length. The latter is saturated at all points by U2 functions,
each taking two particles from similar places in the two different halves
of the cycle. The first term which was ignored in the preceding section,
is obtained from a four particle cycle and two U2 functions, and simple
reasoning shows that its value is:
{[2{h)ex]} (69)
Obviously both kinds of terms (68) and (70) can easily be grouped by
introducing the operator U2 [1 + .rjPex], but we also note that another
modification of l ^ is necessary as well. The reason is that, if we want
to avoid a limitation of the validity of the results to lowest order in the
degree of degeneracy of the system, when calculating the contribution of
any class of diagrams we must include all those that contain additional
arbitrarily long exchange cycles. In other words, as in Sections 3.2 and
4.1, we must incorporate an arbitrarily large number of intermediate U\
functions at all places where they can be inserted into this new class
of diagrams, without changing their topology. Moreover, after inserting
these chains of JJ\ functions at any place between the JJ2 functions, we
must take the sum over the lengths of the chains. But we already know
from Section 3.2 what the effect of this process is: the summation merely
introduces operators [l — nQ^V\\~ • More precisely, since in the present
case this operation has to be carried out twice, once for each particle
contained in the U2 functions, two fractions are introduced. Therefore,
to incorporate all diagrams that we have discussed, we have to replace
I/2 in (68) by the new operator U2 defined as:
Finally we have to sum over all values of k. The situation, then, becomes
reminiscent of that for the ideal gas: because all pairs of particles
124 F. Laloe
appearing in the U2 functions play the same role, it turns out that the
counting factor of the diagrams is simply given by 1// = 2k. This implies
that one again obtains a logarithmic function,! so that the term to be
added to the logarithm of z is:
-Tr{Log[l-e 2 ^l/ 2 (l,2)]}. (72)
f Also, for attractive interactions, the critical values of ft are negative, so that we do not
have to worry about singularities in (71).
J Of course, it is perfectly possible that another condensation process involving more than
two particles will, in turn, take over and quench the condensation of pairs; this possibity
includes ordinary condensation into a liquid or even a solid phase. For a brief discussion
of other condensation processes see next section.
126 F. Laloe
constant and positive, when /? increases the exponential varies rapidly
and always becomes smaller, at some point, than the eigenvalue. A sin-
gularity will then necessarily occur: even weakly attractive potentials
will allow a Bose-Einstein condensation of pairs. We therefore recover
a well-known property of fermions: at sufficiently low temperature,
arbitrarily small attractions destabilize the Fermi sphere, an essential
ingredient of the BCS theory [30]. If the temperature is not very low,
and if there is no well-formed Fermi sphere with a sharp limit (the
usual assumption in the BCS theory), formulas (71) and (72) still apply
without any particular modification; the fractions in (71) automatically
take care of the cross-over between the degenerate and classical regime,
as noted above. We note in passing that, instead of the potential or
of the scattering length which often appears in the theory of pairing
in Fermi systems [31], here the relevant matrix element is that of an
operator which depends both directly on the temperature through the
definition of I/2 and indirectly through the factors [1 — p/] in the def-
inition of U2', moreover it contains all partial scattering waves and
not only the s-wave. In addition, one may use (71) and (72) to dis-
cuss the cross-over between the two regimes where condensation takes
place, either for strongly bound binary molecules, or weakly bound BCS
pairs [32, 33, 34].f
A final remark is that what we have discussed, for simplicity, is the
condensation of bare, non-interacting, pairs. But by introducing the
contribution of U3 operators (and by replacing U\ by U\), it would be
possible to extend the theory in the same spirit as that of Section 4.1 and
define a dressing of the operator £/2 under the effect of the interactions
of pairs with a third partner which would lead to the condensation of
"quasipairs of particles".
f For more specific discussions including films, see Refs. [35] and [36].
Dilute-Degenerate Gases 127
In nuclei, four fermion condensate has been discussed in the context of
the condensation of a particles [37].
But even all these condensation channels do not necessarily exhaust all
the diagrams contained in the partition function. If all sorts of diagrams
can indeed be grouped, as above, in condensation channels where they
generate logarithms, then all the physics is contained in competing Bose-
Einstein condensation processes of composite particles. If some other
diagrams cannot be regrouped in this way, then there will be a channel
for formation of a "non-monochromatic momentum peak", as discussed
in Section 3.3, in competition with the ordinary ("monochromatic")
condensation channels. Of course there is no proof at this stage that such
processes can exist, for a given class of interaction potentials, so that this
remark remains speculative.
Finally, we note that the self-consistent equation (64) which defines the
U operator of the quasiparticles is non-linear, and may contain insta-
bilities if the coupling is strong enough and the temperature sufficiently
low. It may happen that a spontaneous symmetry breaking process, anal-
ogous to ferromagnetism, will allow one to obtain lower values of the
grand potential than the values obtained without this process; this would
correspond to a superfluid instability.
5 Conclusion
In this article, we have used another version of perturbation theory, not
in terms of the interaction potential, but in terms of Ursell operators
which give rise to various contributions in the form of U-C diagrams.
This leads to the exact expression (42) for the grand potential of the
system, expressing it as a sum of various terms arising from these dia-
grams. Each of these terms is obtained as an integral (a trace) over a finite
number of variables. The expression is valid for dilute or dense systems
such as liquids or solids, but for gases it becomes simpler because it can
be truncated conveniently. A feature of the formalism is that the size
of exchange cycles appears explicitly, which provides a natural connec-
tion with the numerical work of Ceperley and Pollock [38], emphasizing
the role of cyclic exchange of all orders in Bose-Einstein condensa-
tion and superfluidity. In our case, the summations over the lengths
of cycles are straightforward, introducing simple functions (fractions)
of the operators. Several mechanisms can lead to a divergence of the
size of the exchange cycles in the relevant diagrams; each of them in-
troduces a separate channel for Bose-Einstein condensation. In other
128 F. Laloe
words, the formalism provides information on various physical pro-
cesses which occur in parallel, such as the condensation of particles
(or quasiparticles), that of pairs, or the contribution of larger clus-
ters, without requiring any ad hoc additional assumption. Refining
the results on a given condensation channel by including interactions
through a renormalization of the operators is also possible, at least
formally, by the introduction of integrals of finite dimension. For in-
stance, if necessary, one could improve the theory of condensation of
pairs by replacing, in the statistical factors appearing in the definition
of I/2, the bare operators U\ by dressed operators U. The method
gives simple formulas that remain valid at both low and high tem-
peratures (degenerate or classical systems); in particular, it includes
automatically all s,p,d,... collision channels. It could be adapted rel-
atively easily to the inclusion of external potentials (inhomogeneous
BEC), since this additional interaction is entirely contained in U\.
We plan to use our method for calculating condensate fractions in
this kind of problem. A more systematic comparison with the re-
sults of the literature is also desirable: energy of the ground state
for bosons [39, 18] and fermions [40, 19], as well as pseudopotential
calculations of the magnetic susceptibility for fermions. Finally, the
study of superfluid gases, especially in the case considered by Siggia and
Ruckenstein where there are two internal states which may condense
simultaneously [41], is another natural domain of application of our
formalism.
References
[1] T.J. Greytak, this volume.
[2] I. Silvera, this volume.
[3] T.J. Greytak and D. Kleppner, in New Trends in Atomic Physics
Vol. II, G. Grynberg and R. Stora, eds. (North-Holland, N.Y, 1984),
p.1127.
[4] Y. Castin, J. Dalibard, and C. Cohen-Tannoudji, this volume.
Dilute-Degenerate Gases 129
[5] J.P. Wolfe, Jia Ling Lin, and D.W. Snoke, this volume.
[6] A. Mysyrowicz, this volume.
[7] L.V. Keldysh, this volume.
[8] O. Penrose and L. Onsager, Phys. Rev. 104, 576 (1956).
[9] RE. Sokol, this volume.
[10] N. Bogoliubov, Journal of Physics USSR 11, 23 (1947).
[11] Kerson Huang, Statistical Mechanics, (Wiley, N.Y., 1963) Section
13.3; sec. ed. (1987), Section 10.5.
[12] K. Huang and C.N. Yang, Phys. Rev. 105, 767 (1957); K. Huang,
C.N. Yang and J.M. Luttinger, Phys. Rev. 105, 776 (1956); T.D. Lee,
K. Huang and C.N. Yang, Phys. Rev. 106, 1135 (1957).
[13] T.D. Lee and C.N. Yang, Phys. Rev. 112, 1419 (1958); 113, 1406
(1959).
[14] T.D. Lee and C.N. Yang, Phys. Rev. 113, 1165 (1959); 116, 25
(1959); 117, 12, 22 and 897 (1960).
[15] R.K. Pathria, Statistical Mechanics, (Pergamon, Oxford, 1972), Sec-
tion 9.7.
[16] K. Huang, loc. cit, Chapter 14; second edition (1987), Chapter 10.
[17] R.K. Pathria, Ref.[15], Section 9.6.
[18] S.T. Belyaev, Soviet Phys. JETP 34, 289 (1958).
[19] V.M. Galitskii, Soviet Phys. JETP 34, 104 (1958).
[20] E.P. Gross, Journ. Math. Phys. 4, 195 (1963).
[21] F. Laloe, J. Physique, in preparation
[22] M. Pinard and F. Laloe, J. Physique 41, 769 (1980).
[23] M. Pinard and F. Laloe, J. Physique 41, 799 (1980).
[24] C. Lhuillier et F. Laloe, J. Physique 43, 197 and 225 (1982).
[25] PJ. Nacher, G. Tastevin and F. Laloe, J. Physique I 1, 181 (1991).
[26] K. Huang, in Studies in Statistical Mechanics, Vol. II, J. De Boer
and G.E. Uhlenbeck, eds. (North-Holland, Amsterdam, 1964).
[27] C.N. Yang and T.D. Lee, Phys. Rev. 87, 404 (1952).
[28] K. Huang, C.N. Yang and J.M. Luttinger, Phys. Rev. 105, 776
(1957), Section 6.
[29] C.N. Yang, Rev. Mod. Phys. 34, 694 (1962).
[30] J. Bardeen, L.N. Cooper and J.R. Schrieffer, Phys. Rev. 108, 1175
(1957).
[31] E.M. Lifshitz and L.R Pitaevskii, Statistical Physics, Vol. II (Perga-
mon, Oxford, 1980), Sections 39-40.
130 F. Laloe
[32] A.J. Leggett, in Modern Trends in Condensed Matter Physics, Pro-
ceedings 16th Karpacz Winter School, A. Pekalski and J. Przystawa,
eds. (Springer, Berlin, 1980); AJ. Leggett, J. Physique coll. C7, 19
(1980).
[33] See review by M. Randeria, this volume.
[34] P. Nozieres and S. Schmitt-Rink, J. Low Temp. Phys. 59, 195 (1985).
[35] D.M. Eagles, Phys. Rev. 186, 456 (1969).
[36] K. Miyake, Progr. Theor. Phys. 69, 1794 (1983).
[37] M. Apostol, Phys. Lett. 110A, 141 (1985).
[38] D.M. Ceperley and E.L. Pollock, Phys. Rev. Lett. 56, 351 (1986);
Phys. Rev. B36, 8343 (1987); E.L. Pollock and D.M. Ceperley, Phys.
Rev. B 39, 2084 (1989).
[39] Lifshitz and Pitaevskii, Ref.[31], Section 25.
[40] Lifshitz and Pitaevskii, Ref.[31], Section 6.
[41] E.D. Siggia and A.E. Ruckenstein, Phys.Rev.Lett. 44, 1423 (1980);
J. Physique 40, C7 15 (1980); Phys. Rev. B 23, 3580 (1981).
Prospects for Bose-Einstein Condensation
in Magnetically Trapped Atomic
Hydrogen
Thomas J. Greytak
Physics Department
Massachusetts Institute of Technology
Cambridge, MA 02139
USA
Abstract
Atomic hydrogen would be an ideal substance in which to observe Bose-
Einstein condensation. Recombination of the atoms into molecules can be
slowed dramatically by polarizing the electronic spins in a high magnetic
field at low temperature, but achieving the necessary density in samples
confined by material walls has not been possible. Evaporative cooling of
magnetically trapped spin-polarized hydrogen is a more promising approach.
It has already produced temperatures as low as 100 /iK (well below the
laser cooling limit) at a density of 8 x 1013 cm" 3 . At this density, the
transition temperature is 30 //K. Reaching the transition by evaporative
cooling appears to be possible. Laser techniques are being developed to
detect the transition and to study the properties of the resulting gas.
1 Introduction
h2 / n y/3
2nkBM V2.612/
/ n \2/3
2O (1)
U.965xlO J '
131
132 T. J. Greytak
and pointed out that the gas would be a mixture of two of the hyper-
fine states of hydrogen. The idea was raised again by Stwalley and
Nosanow [2] in 1976. By this time more complete ground state energy
calculations had been done confirming the absence of a liquid phase
for atomic hydrogen. They pointed out that the presence of the in-
teraction between the atoms could lead to the Bose condensed state
being a superfluid as well. The weakness of the interaction and the
ability to change the density would allow a detailed comparison be-
tween theory and experiment. Although Stwalley and Nosanow saw
BEC as the ultimate long term goal of spin-polarized atomic hydrogen
research, they also expected that "these systems should yield exciting
and fundamental new information about the properties of quantum
systems".
The ultimate goal of BEC in hydrogen has yet to be realized, though
we seem to be quite close at the current time. On the other hand,
since spin-polarized atomic hydrogen was first created in the labora-
tory by Silvera and Walraven [3], the prediction of exciting results
from the "new" quantum system has certainly been fulfilled. Once
the atomic hydrogen has been electron spin polarized, it spontaneously
develops nuclear polarization [4, 5], The doubly polarized gas was
found to support sharply defined nuclear spin waves [6-8], even though
the mean separation between the atoms is much larger than their De-
Broglie wave length. The doubly polarized gas, when compressed to
higher densities [9, 10], became a model system for the study of three-
body recombination processes [11]. Cryogenic atomic hydrogen masers
[12-14] offer the the possibility of significant increases in frequency
stability over similar room temperature devices [15]. Most recently,
the cold hydrogen atoms have been used to study atom-surface inter-
actions, in particular the sticking probability of very cold atoms on
well-understood surfaces [16-18]. These aspects of spin-polarized atomic
hydrogen have been the subject of two extensive reviews [19, 20]. In
this article I will concentrate on the possibilities for attaining BEC.
I will review some of the early prospects and explain why they did
not come to fruition. I will discuss magnetic trapping and evapora-
tive cooling of the hydrogen, which seems to be the most promising
avenue toward BEC at present. And finally, I will examine what the
condensed state might look like in the trapped gas, and how it could be
studied.
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 133
<D
C
-0.2 -
-0.4 -
2 Early Approaches
Figure 1 shows the splitting of the ground state of atomic hydrogen due
to the hyperfine interaction. The two lowest energy states, labeled a and
b, are called 'high field seekers' since they are drawn into a region of high
magnetic field. Similarly, the two highest energy states, c and d, are 'low
field seekers'. All four states are created in roughly equal amounts when
molecular hydrogen is dissociated in an RF discharge. In the simplest
experiments the atomic hydrogen develops electron polarization when the
atoms impinge on a region of high magnetic field (typically the center
of a superconducting magnet at about 8 Tesla) and low temperature
(typically 300 mK obtained by a dilution refrigerator). Atoms in the c
and d states are excluded from the high field region; those in the a and b
134 T. J. Greytak
(a)
Fig. 2. Sample chambers for high field seekers used by the MIT group, (a) For
simple fills. The inner diameter is 1 cm.
states enter, lose their excess energy through wall collisions, and become
trapped.
It is unfortunate that Maxwell's equations do not allow a maximum in
the magnitude of the magnetic field in a source free region. Otherwise,
the high field seekers could be held in a pure magnetic trap. In practice
the atoms are usually confined radially and from below by physical walls,
as is shown in Fig. 2(a). One could say that the magnetic field acts as a
cork in an otherwise open bottle.
The electron-polarized gas of a and b atoms is not stable against
recombination. The finite admixture of the 'wrong' direction of elec-
tron spin in the a state due to hyperfine mixing allows the reactions
a + b —• ortho H2 and a + a —• para H2. The recombination of
two hydrogen atoms releases 4.476 eV of energy. Initially most of this
energy goes into rotation and vibration of the molecule, leaving less
than 1% for translational motion [11]. Since a two-body collision can-
not conserve momentum while simultaneously releasing kinetic energy,
these reactions require the presence of a third body. The third body
could be another hydrogen atom, but at the densities achieved in simple
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 135
-Glass plate
\ Z:
Pressure
transducer
Film Sample
thermometers ^/chamber
f !spacer
Indium O-ring
Silver sinter-
-Piston
(b)
Fig. 2. (b) For compression experiments. The hydrogen is compressed into the
region above the sintered silver, 90 /urn high and 0.8 cm in diameter.
fill experiments the reaction takes place while the atoms are adsorbed
on the walls of the cell with the wall acting as the third body. Re-
combination is slowed, but not eliminated, by coating the walls of the
cell with superfluid helium to minimize the van der Waals attraction
between the atoms and the surface. The result of these reactions is
to remove all of the a atoms while leaving a finite fraction of the b
atoms. The resulting gas of b atoms is nuclear as well as electron-spin-
polarized.
Although the gas of pure b atoms is stable against recombination,
nuclear relaxation can occur during collisions [4], causing a transition
to the a state. The nuclear relaxation is driven by the fluctuating mag-
136 T. J. Greytak
hyperfine state b
number of atoms 2 x 1016
density 4.5 x 1018 cm"3
temperature 550 mK
lifetime 5 sec
nc at this temperature 2 x 1020 cm"3
Tc at this density 44 mK
Source solenoid
Confined atoms Quadrupole magnet
Hyperfine resonator
,1,,,;;;;,,,;,;,,,,,,r?
Source
Loss by evaporation
E/atom » <E>
Causes cool
Hotest Gas cools,
atoms settles deeper
leave into trap
Loss by spin-relaxation
E / atom < <E> Density stays about constant
Causes heating
the trap center. The atomic hydrogen source is located in a high field
solenoid.
Figure 4 illustrates the trapping and cooling process. In Fig. 4(a)
two important energies are introduced. The potential energy difference
between the center of the trap and the wall of the cell is represented by
£wau. The radial trapping fields are proportional to £waii- In practice the
lowest barrier to escape from the trap, ^threshold, occurs on axis at the
lower pinch solenoid, ^threshold governs the evaporative cooling process.
Atoms which manage to reach a helium covered surface, either by going
radially out to the wall or (much more likely) by escaping over the
threshold and entering the bottom part of the cell, can be adsorbed and
recombine into molecules.
Collisions between the atoms are essential to the loading of the trap.
Otherwise, atoms from the source would accelerate into the trapping
region, move across unimpeded, ride up the far side of the potential and
leave the trap over ^threshold- Radially, the atoms would bounce back
and forth between the walls. It is only through atom-atom collisions
that a fraction of the atoms lose enough energy to become trapped
near the center of the cell. Soon all atoms with energies greater than
140 T. J. Greytak
Atom
source
Magnet
system
Bolometer
Retro mirror
O-k^m MicroChannel
plate
Fig. 5. The magnetic trap in current use at MIT.
gas. A scale drawing of the new trap is shown in Fig. 5. I will discuss its
operation and capabilities in some detail since this trap, or one like it,
will be used to try to achieve BEC in atomic hydrogen.
The loading of the trap and the subsequent development of electronic
and nuclear polarization in the gas can be understood with the help of
the simplified sketches in Fig. 6. Molecular hydrogen is condensed as a
solid on the walls of the source region during the initial cool down of
142 T. J. Greytak
J LHe H He
He He H
H H H
H2 H2 H He
He He
He He H H
He He H H
He H
He H2
He H H
H2 He H HeH H
HeH H
He He
H2 He H
H H
He H2 H He H He
n r
H2 He H He
T'r H
H
He
Hc
H He
H
He
H
H
He
He
J
H
He
He
H
J L
He He
J L
He He
J L
He He
He He He He He He
He He He He He He
H
He He He He
i
He d He
dC
d
He cJ
He He He He d d
He
c1
dd
H
H d! d
He He He He He He
H
H d d
H c
He H He He He He He
He He He He He He
He He He He He He
He He He He He He He He He
Trap after helium Trap after high field seekers Trap after spin-exchang
recondenses return to the source eliminates c states
Fig. 6. Steps leading to a doubly polarized gas of low field seeking hydrogen
atoms in the MIT trap.
hyperfine state d
number of atoms 2 x 1014
density on axis 2 x 1013 cm"3
temperature 50 mK
lifetime 20018sec 3
nc at this temperature 6 x 10 cm"
Tc at this density 12 ^K
the magnetic resonance signal [27, 28] or, more easily, from the energy
liberated when the released atoms recombine [30]. After nuclear po-
larization has been established, the time evolution of the total number
of atoms in the trap, N(t), is governed by spin relaxation and depends
on the trap shape, the electronic spin relaxation rate constant g, and
the temperature of the gas. Studies of N(t) for isolated atoms [27]
(where T was estimated) and for atoms in weak thermal contact with
the walls [29] (where T was known) confirmed the values of g calculated
theoretically [31, 32]. Once g was established, the temperature of the
gas and its density could be determined indirectly from an analysis of
N(t) [28].
A more direct method of measuring the gas temperature, developed
by Doyle [30], is illustrated in Fig. 7. A highly sensitive bolometer is
used to measure the flux of atoms from the trap as ^threshold is ramped
down to zero at a uniform rate. If the time necessary to drop the field
is short compared to the thermal equilibration time in the gas, the flux
versus time gives the number density versus energy, n(E). The energy
distribution n(E) is important in itself; for example it can be used to study
the approach to equilibrium during and after evaporative cooling. The
temperature of the gas is determined by comparing n(E) to the calculated
distribution expected for the known magnetic field profile. Figure 8 shows
n(E) measured after different degrees of evaporative cooling in the MIT
trap [17]. The lowest temperature achieved in that experiment, 100 /xK,
is well below the recoil limit for Lyman-a laser cooling of hydrogen,
T = (hv)2/2mc2kB « 650 JJK. Some of the features of this gas are given
in Table 4.
Other properties of this coldest hydrogen gas may be of interest. The
mean thermal speed, v = ^%kBT/nM, is 145 cm/s. The gravitational
height, kBT/Mg, is 8.4 cm and is beginning to play a role in the vertical
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 145
Fig. 7. The energy distribution in the gas, n(£), can be determined by measuring
the flux of atoms out of the trap as the threshold for escape is lowered quickly
to zero.
12 16
Energy [mK]
200 300
Energy
Fig. 8. Energy distributions for evaporatively cooled atomic hydrogen. The solid
curves are the calculated distributions for atoms at (a) 1.1 mK, (b) 190 //K, and
(c) 100 fiK. The dashed curves in (c) are for temperatures 30% higher and lower
than the best fit value. The finite atom signals from E < 0 are due to a small
axial field inhomogeneity ( < 1 G) that slightly deforms the bottom of the trap.
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 147
hyperfine state d
number of atoms 1011
3 x 13
density on axis 8 x 10 cm"3
temperature 100 ^K
lifetime 50 sec
nc at this temperature 5 x 1014 cm"3
Tc at this density 30 ^K
for a uniform gas at a density of 8 x 1013 cm 3 the mean free path would
be 12 cm and the mean free time, X/v, would be 83 ms. It is clear, then,
that the trapped atoms traverse the confining potential hundreds of times
between collisions.
As Table 4 indicates, one would have to cool this gas by only a little
more than a factor of three in order to attain BEC. Is this feasible?
For evaporative cooling to be effective the thermalization rate for the
gas (proportional to the collision rate) must be greater than the spin
relaxation rate. Both of these rates are proportional to the density,
but the collision rate is also proportional to the square root of the
temperature whereas the relaxation rate is independent of temperature
for low temperatures and fields. Therefore the efficiency of evaporative
cooling decreases as the temperature is lowered [26]. Nevertheless, model
calculations [34] indicate that evaporative cooling using the current MIT
trap can cross the phase boundary for BEC at densities between about
5 x 1013 and 5 x 1014 cm" 3 . In order to observe the phase transition, it
would be very useful to have a non-destructive means of studying the gas.
The following section discusses the expected properties of the hydrogen
gas as it undergoes BEC in the trap and a possible unambiguous way of
observing the phase transition.
nc(T) = 2.612A(T)-3
= 4.965 x 1020T3/2, (3)
i/ = (Snan)-1/2
= 2.35 x 10 3 ^p. (4)
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 149
The interaction strength Vo has the units of energy times volume
u = ^nan)1'2. (6)
nVo ~ V 2 u
1.75 x 1010
There has long been concern about the time necessary to nucleate the
Bose-Einstein phase transition in a low density gas once the critical
density or temperature has been reached. Recent theoretical calculations
by Stoof [39] and Kagan, Svistunov and Shlyapnikov [40] indicate that
the time is short compared to the expected lifetime of the trapped gas.
The details of the two theories differ (see the articles by Kagan and Stoof
in this book.) However, there seems to be agreement that an observable
peaking of the energy density n{E) in a group of states close to E = 0
will occur on a timescale of the order of TC. For a particle density of
n = 8 x 1013 cm"3, the characteristic time is TC = 0.2 ms.
Let us consider a specific example of a configuration in which BEC
might occur in the trapped hydrogen. By the time the magnetic fields
have been reduced to relatively low values during the evaporative cooling
process, the potential near the center can be approximated by a parabolic
minimum. There is a great deal of freedom in choosing the exact shape
of the potential in which the coldest atoms will reside. One possible
potential which is both attainable and convenient is
U(r,z)/kB = 1.4 x 10~2r2 + 3.0 x 10"5z2, (8)
150 T. J. Greytak
z=0
where the radial (r) and axial (z) coordinates are expressed in centimeters
and U/ks is expressed in units of degrees Kelvin. In an actual experiment
the trap shape, temperature and number of atoms will all be changing
simultaneously. I will consider here a simpler hypothetical situation where
the trap shape and the temperature are held fixed and the total number
of atoms in the trap is slowly increased. I will take the temperature to be
30 /iK. At this temperature the critical density is nc(T) = 8 x 1013 cm" 3 .
When N is so small that the density at the center of the trap is much
less than nc(T)9 the density profile is proportional to e~u^kBT', as in the
case of a classical gas. The surface on which the density has fallen to e~ l
of its maximum value is ellipsoid of revolution extending to r = 460 /am
at z = 0 and z = ±10000 fim on axis (22 times longer than it is wide).
This is shown in Fig. 9. As N increases the density distribution develops
more weight at the low values of U than would be expected from the
classical form. Critical density is reached at the center of the trap when
N = 4.32 x 1011. Consider the various distance scales. The thermal
DeBroglie wavelength A(T) = 0.32 fim at this temperature. The healing
length Y\ = 2.6 /mi at this density. If there were no interactions between
the particles, the atoms would begin to condense into the single particle
ground state associated with this harmonic potential. That wave function
would have a radial extent of about 6 jum, too narrow to be resolved
in Fig. 9. Because of the repulsive interactions, the actual condensate is
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 151
spread over a wider region. For example, when N reaches 4.55 x 1011
the fraction of atoms in the condensate is 2.5% and the phase boundary
extends out to r = 63 irni.
Once a condensed state has formed in the gas, the density at the center
of the trap increases rapidly above nc(T) as more atoms are added. Since
the atom loss rate due to electronic relaxation is proportional to the
square of the density, the gas should begin to decay more rapidly once
BEC has been established. Hijmans et al. [41] have suggested that the
density at the trap center builds up so quickly with increasing condensate
fraction that the resulting rapid relaxation and subsequent recombination
can be likened to an 'explosion' (see the contribution by Hijmans et al. in
this book.) In order to study this possibility for the specific trap discussed
above it is necessary to compute the complete density distribution, n(r).
A simple thermodynamic approach to calculating n(r) has been intro-
duced by Oliva [42]. In this approach the chemical potential ja is taken
to be constant throughout the system. At each point \i is assumed to be
the sum of the 'internal chemical potential' of the gas, /xo(n(r)), and the
potential energy at that location:
/i = /io(n(r)) + l7(r). (9)
The internal chemical potential is simply the chemical potential of a
homogeneous gas of the same density in the absence of an applied
potential. For the weakly interacting Bose gas|
fio(n(r)) = V0(n + nc(T)), n>nc(T), (10)
= 2nVo+kBT\nz, n < nc(T). (11)
In the normal state the expression for /zo(n(r)) involves the fugacity,
z = ^ O / / C B T , which must be determined from the relation
where gs/2(z) is the familiar function from the theory of the ideal Bose
gas. To find n(r) one first chooses a value for \x. Then at each point in the
gas \x0 is found from Eq. (9) and either Eq. (11) or Eq. (12) is inverted to
find the density at that point. Finally, one integrates n(r) over the trap to
find the total number of atoms N corresponding to that value of \i. Since
N is a monotonically increasing function of /*, one can watch n(r) evolve
t Expressions for most of the thermodynamic variables in a weakly interacting Bose gas
are given in the review by Greytak and Kleppner [19]. The chemical potential is most
easily found from the fact that it is the Gibbs function per particle: \i0 = G/N =
(E-TS+PV)/N.
152 T. J. Greytak
as the trap is filled by systematically increasing the values of \x used in
the procedure. This 'local density approximation' approach to finding
n(r) should be accurate except in those regions where n(r) is changing on
a scale comparable to the healing length.
The expression for the internal chemical potential in the condensed
state, Eq. (11), can easily be inverted and used with Eq. (9) to give an
analytic expression for the density:
0.04
Fig. 11. Behavior of critical parameters during Bose condensation in the trapping
potential of Fig. 9.
results of the model to the observed lineshapes. The same source has
been used to demonstrate Doppler cooling and light-induced evaporation
of trapped hydrogen [47]. In the latter case optical absorption is used
to remove high energy atoms from the trap by turning them into high
field seekers. Using this technique, they were able to cool the trapped
hydrogen to 3 mK.
Jon Sandberg at MIT developed a different optical probe, based on
two-photon absorption. The process is illustrated in Fig. 12. A continuous
source of radiation at one half the 1S-2S frequency is focused to a thin
line in the trapped gas (see also Fig. 9). The light is reflected back along
the same path by a mirror located at low temperature. Some atoms
undergo a two-photon transition from the IS to the 2S state. The 2S
state is metastable, so a small electric field is applied to mix the 2S
and the 2P states. The atoms then decay rapidly from the 2P state by
emitting a Lyman-a photon. The number of Lyman-a photons detected
by a micro-channel plate at low temperature is directly proportional
to the number of hydrogen atoms along the light beam path. The
wavelength dependence of the detector sensitivity strongly discriminates
against light from the source itself. To further decrease the sensitivity
to the illuminating radiation, the source can be chopped and the Stark
mixing of the 2S and 2P states delayed until the 'off' interval.
One way of using the two-photon absorption to study the gas is to
map out the radial density profile by translating the beam across the gas
parallel to the axis (in practice, it may be simpler to move the gas across
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 155
i»
Laser beam ' l*
Mirror
s
Trapped gas
Doppler shifted
"k + ~k
-~k + (-~k)
Doppler free
"k + (-~k)
Fig. 12. Two-photon absorption in a cold hydrogen gas.
the beam). If the trap potential is known, the radial extent of n(r) gives
a measure of the temperature. BEC could be identified by the sudden
appearance of a sharp peak in the density at the center of the trap, as in
Fig. 10.
A more direct signature of the phase transition can be obtained by
measuring the spectrum of the absorption as the frequency of the source
is scanned about the atomic frequency. Figure 12 shows that there are
two types of processes which contribute to the two-photon absorption.
If one photon is taken from each of the two counter propagating waves,
there will be no first order Doppler shift associated with the atom's
motion. This process gives rise to an extremely narrow absorption line
(of the order of 3 KHz wide due primarily to the time of flight of an atom
across the beam [48]), centered at half the atomic frequency v o /2, with
a strength proportional to the total number of atoms illuminated. On
the other hand, if the two absorbed photons were traveling in the same
direction, there would be a Doppler shift of the absorption frequency
' «N
T>T C
RECOIL SHIFT
the atom's momentum along the axis. The familiar p dependent Doppler
term (which mirrors the probability distribution for a given component
of the atom's momentum) is displaced to higher frequencies by a recoil
term. In most gases the recoil shift is negligible compared to the width
of the Doppler spectrum; but in trapped hydrogen, because of the light
mass and low temperature, the width and the shift are comparable. This
is shown schematically in Fig. 13.
By sweeping the frequency of the source one can measure the momen-
tum distribution and, of course, deduce the temperature. The signal will
be quite weak compared to the Doppler free case, however, since this
part of the absorption spectrum is much broader than the line width
of the source. The situation changes below the transition when a finite
fraction of the atoms come almost to rest. In this case a sharp feature will
appear in the absorption centered at the recoil shift with a strength pro-
portional to the condensate fraction (see Fig. 13). Kagan, Svistunov and
Shlyapnikov [40] estimate that the states which develop high occupation
below the transition are within a range A£ ~ nV0 of E = 0. This would
give rise to a frequency spread of the order of one KHz in the recoil
shifted condensate line, which is less than the time of flight broadening
expected for the Doppler free component. These considerations have
led Sandberg [48] to suggest that measuring the two-photon absorption
spectrum may be the simplest and least ambiguous way of detecting BEC
in trapped hydrogen.
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 157
The continuous wave laser source for the two-photon absorption has
been completed. It produces 20 mW of tunable light at X = 243 nm
with a spectral width of about 2 KHz. Work is now underway to see the
absorption signal in the cold hydrogen.
References
[I] C. E. Hecht, Physica 25, 1159 (1959).
[2] W. C. Stwalley and L. H. Nosanow, Phys. Rev. Lett. 36, 910 (1976).
[3] I. F. Silvera and J. T. M. Walraven, Phys. Rev. Lett. 44, 164 (1980).
[4] B. W. Statt and A. J. Berlinsky, Phys. Rev. Lett. 45, 2105, (1980).
[5] R. W. Cline, T. J. Greytak and D. Kleppner, Phys. Rev. Lett. 47,
1195 (1981).
[6] E. P. Bashkin, Pis'ma Zh. Eksp. Teor. Fiz. 33 , 11 (1981) [Soviet
Physics JETP Lett. 33, 8 (1981)].
[7] C. Lhuillier and F. Laloe, J. Phys. (Paris) 43, 197, 225 (1982).
[8] B. R. Johnson, J. S. Denker, N. Bigelow, L. P. Levy, J. H. Freed and
D. M. Lee, Phys. Rev. Lett. 52, 1508 (1984).
[9] D. A. Bell, H. F. Hess, G. P. Kochanski, S. Buchman, L. Pollack,
Y. M. Xiao, D. Kleppner and T. J. Greytak, Phys. Rev. B 34, 7670
(1986).
[10] R. Sprik, J. T. M. Walraven and I. F. Silvera, Phys. Rev. B 32, 5668
(1985).
[II] Yu. Kagan, I. A. Vartanyantz and G. V. Shlyapnikov, Sov. Phys.
JETP 54, 590 (1981).
[12] H. F. Hess, G. P. Kochanski, J. M. Doyle, T. J. Greytak and D.
Kleppner, Phys. Rev. A 34 ,1602 (1986).
158 T. J. Greytak
[13] M. D. Hiirlimann, W. N. Hardy, A. J. Berlinsky and R. W. dine,
Phys. Rev. A 34, 1605 (1986).
[14] R. L. Walsworth Jr, I. F. Silvera, H. P. Godfried, C. C. Agosta, R.
F. C. Vessot and E. M. Mattison, Phys. Rev. A 34, 2550, (1986).
[15] A. J. Berlinsky and W. N. Hardy, 13th Annual Precision Time and
Time Interval Applications and Planning Meeting, 1981, NASA
Conference Publication 220, p. 547.
[16] J. J. Berkhout, E. J. Wolters, R. van Roijen and J. T. M. Walraven,
Phys. Rev. Lett. 57, 2387 (1986).
[17] J. M. Doyle, J. C. Sandberg, I. A. Yu, C. L. Cesar, D. Kleppner and
T. J. Greytak, Phys. Rev. Lett. 67, 603 (1991).
[18] I. A. Yu, J. M. Doyle, J. C. Sandberg, C. L. Cesar, D. Kleppner and
T. J. Greytak, Phys. Rev. Lett. 71, 1589 (1993).
[19] T. J. Greytak and D. Kleppner, in New Trends in Atomic Physics, G.
Grynberg and R. Stora, eds. (North-Holland, Amsterdam, 1984).
[20] I. F. Silvera and J. T. M. Walraven, Spin polarized atomic hydrogen,
in Progress in Low Temperature Physics, D.F. Brewer, ed. (North-
Holland, Amsterdam, 1986), Vol. X, p. 139.
[21] T Tommila, E. Tjukanov, M. Krusius and S. Jaakkola, Phys. Rev.
B 36, 6837 (1987).
[22] H. T. C. Stoof, L. P. H. de Goey, B. J. Verhaar and W. Glockle,
Phys. Rev. B 38, 11221 (1988).
[23] J. D. Gillaspy, I. F. Silvera and J. S. Brooks, Phys. Rev. B 40, 210
(1989).
[24] Yu. Kagan and G. V. Shlyapnikov, Phys. Lett. A 130, 483, (1988).
[25] E. Tjukanov, A. Ya. Katunin, A. I. Safonov, P. Arvela, M. Karhunen,
B. G. Lazarev, G. V. Shlyapnikov, 1.1. Lukashevich and S. Jaakkola,
Physica B 178, 129 (1992).
[26] H. F. Hess, Phys. Rev. B 34, 3476 (1986).
[27] H. F. Hess, G. P. Kochanski, J. M. Doyle, N. Masuhara, D. Kleppner
and T. J. Greytak, Phys. Rev. Lett. 59, 672 (1987).
[28] N. Masuhara, J. M. Doyle, J. Sandberg, D. Kleppner, T. J. Greytak,
H. F. Hess and G. P. Kochanski, Phys. Rev. Lett. 61, 935 (1988).
[29] R. van Roijen, J. J. Berkhout, S. Jaakkola and J. T. M. Walraven,
Phys. Rev. Lett. 61, 931 (1988).
[30] J. M. Doyle, J. C. Sandberg, N. Masuhara, I. A. Yu, D. Kleppner
and T. J. Greytak, J. Opt. Soc. Am. B 6, 2244 (1989).
[31] A. Lagendijk, I. F. Silvera and B. J. Verhaar, Phys. Rev. B 33, 626
(1986).
Prospects for BEC in Magnetically Trapped Atomic Hydrogen 159
[32] H. T. C. Stoof, J. M. V. A. Koelman and B. J. Verhaar, Phys. Rev.
B 38, 4688 (1988).
[33] D. G. Friend and R. D. Etters, J. Low Temp. Phys. 39, 409 (1980).
[34] J. M. Doyle, Ph.D. thesis, MIT (1991).
[35] V. Bagnato, D. E. Pritchard and D. Kleppner, Phys. Rev. A 35, 4354
(1987).
[36] K. Huang, Studies in Statistical Mechanics II, J. DeBoer and G. E.
Uhlenbeck, eds. (North-Holland, Amsterdam, 1964).
[37] W. Kolos and L. Wolniewicz, J. Chem. Phys. 43, 2429 (1965), and
Chem. Phys. Lett. 24, 457 (1974).
[38] E. P. Gross, J. Math. Phys. 4, 195 (1963).
[39] H. T. C. Stoof, Phys. Rev. Lett. 66, 3148 (1991).
[40] Yu. M. Kagan, B. V. Svistunov and G. V. Shlyapnikov, Sov. Phys.
JETP 75, 387 (1992).
[41] T. W. Hijmans, Yu. Kagan, G. V. Shlyapnikov and J. T M. Walraven,
contribution to this volume.
[42] J. Oliva, Phys. Rev. B 39, 4197 (1989).
[43] V. V. Goldman, I. F. Silvera and A. J. Leggett, Phys. Rev. B 24, 2870
(1981).
[44] C. A. Condat and R. A. Guyer, Phys. Rev. B 24, 2874 (1981).
[45] D. A. Huse and E. D. Siggia, J. Low Temp. Phys. 46, 137 (1982).
[46] O. J. Luiten, H. G. C. Werij, I. D. Setija, M. W. Reynolds, T. W.
Hijmans and J. T. M. Walraven, Phys. Rev. Lett. 70, 544 (1993).
[47] I. D. Setija, H. G. C. Werij, O. J. Luiten, M. W. Reynolds, T. W.
Hijmans and J. T. M. Walraven, Phys. Rev. Lett. 70, 2257 (1993).
[48] J. C. Sandberg, Ph.D. thesis, MIT (1993).
8
Spin-Polarized Hydrogen: Prospects for
Bose-Einstein Condensation and
Two-Dimensional Superfluidity
Isaac F. Silvera
Lyman Laboratory of Physics
Harvard University
Cambridge MA 02138
USA
Abstract
1 Introduction
160
Spin-Polarized Hydrogen: BEC and 2D Superfluidity 161
The critical temperature for BEC is given by
Tc = 331(h2/mkB)n2/3 (1)
When evaluated at Tc = 100 mK, n has a value of 1.57 x 1019/cm3.
Hydrogen has been cooled to such temperatures but not at this density.
The difficulty in achieving the conditions for BEC is the rapid intrinsic
recombination rate to molecular hydrogen. This has two effects: (i) it
limits the density, as the recombination rate is proportional to n3 in the
gas phase, and (ii) the 4.6 eV per recombining pair is usually dissipated
in the gas and can result in severe overheating of the sample, preventing
the attainment of Tc.
Hydrogen was initially stabilized in a cell with a thin film of superfluid
4
He covering the copper walls of the container. Hydrogen atoms reside
in states in the gas phase or states on the surface, and the ratio depends
exponentially on the adsorption energy. Helium presents a surface to the
hydrogen atoms with a very small adsorption energy, £a//c# « IK. In the
high temperature limit the surface coverage o of hydrogen is given by
(T = nkexp(sa/kBT), (2)
where k is the thermal de Broglie wavelength. From this we see that as n
increases or T decreases, the coverage builds up. The build up of surface
density can be very detrimental, as on the surface recombination can take
place when two hydrogen atoms collide in the presence of a helium atom.
The rate for this process is proportional to a2. At very high density the
surface recombination is dominated by three-body hydrogen collisions,
proportional to a3; ultimately two- and three-body recombination can
prevent the achievement of low temperatures in a cell or the build up of
gas phase density at low temperatures. Due to this problem some effort
has been expended to use magnetic traps to isolate hydrogen from the
container walls and completely eliminate wall recombination.
A more precise description of the recombination problem shows that
the rates depend on the hyperfine states of the colliding atoms. Due to
the spin 1/2 of the electron and spin 1/2 of the proton, there are four
hyperfine states, delineated a, b, c and d in order of increasing energy,
as shown in Fig. 1. In a high magnetic field the a and b states are
predominantly electron spin-down states, while the c and d states are
spin-up states. The recombination rate constants depend sensitively on
the hyperfine states of the atoms. For example, if two of the atoms in a
recombination collision are in one of the electron spin-up states and one
of the spin-down states, we have the situation for unpolarized hydrogen
162 /. F. Silvera
and the recombination rate constant is very large. The rate constant is
reduced by several orders of magnitude when all atoms are in either the
up-spin states or down-spin states, i.e. spin-polarized. The a and c states
have admixtures of spin-up and spin-down. This admixture becomes very
small in a large magneticfield,which enhances the stability. However, the
b and d states are pure electron spin states and gases made of either one
of these states have still higher stability, as a recombination channel is
closed. Thus, the stability of the gas can be greatly increased by placing
all of the atoms in hyperfine states b or d.
Two-body collisions can also result in relaxation amongst the hyperfine
states. This can have beneficial or detrimental effects. For example,
in the static magnetic trap, the gas very rapidly relaxes to a single
hyperfine state, d, due to spin-exchange processes. This state is very
stable. Ultimately, however, as the conditions for BEC are approached,
magnetic dipolar relaxation dominates the behavior [3]. The d-state
atoms relax to the spin-down states which are ejected from the trap.
These processes have prevented the attainment of BEC in the static trap,
as will be discussed elsewhere in this volume [4].
Spin-polarized hydrogen also exists as an almost two-dimensional gas
when condensed on the surface of helium. The atoms are bound to
the 4He surface by an adsorption energy sa/kB « IK and are in two-
dimensional free particle momentum states; in thermodynamic equilib-
rium the coverage a is given by Eq. (2). At high densities this expression
Spin-Polarized Hydrogen: BEC and 2D Superfluidity 163
is modified according to the Bose distribution; due to effective repulsive
interactions between hydrogen atoms on the surface, the coverage satu-
rates at crsat « l x 10 14/cm2. The surface gas of hydrogen is expected
to demonstrate the behavior of a degenerate Bose gas. Although it can
be shown rigorously that Bose- Einstein condensation does not exist in
two dimensions, it is believed that a quasi-condensate with short-range
correlations will exist and the correlation length rapidly becomes the
size of a laboratory sample as temperature is reduced below the critical
value. The critical temperature for the Kosterlitz-Thouless transition to
two-dimensional superfluidity is
T]D = nh2as/2kBm , (3)
where as ~ a is the superfluid fraction [5]. For GS = <7sat, the saturation
coverage, the calculated critical temperature is 756 mK. It is believed that
coverages of hydrogen in this range can be achieved. Thus, it is quite
realistic to seek the KT transition in hydrogen. Until recently most efforts
have concentrated on achieving BEC in three dimensions; however, it
now seems that it may be easier to deal with the recombination problems
in two dimensions, so new efforts are underway.
In the preceding we have given a brief introduction to spin-polarized
hydrogen and Bose condensation. In the remainder of this paper we shall
concentrate the discussion on two promising approaches: the microwave
trap and two-dimensional quasi-BEC. There are a number of other
interesting ideas for BEC which have been reviewed recently by Silvera [6]
and Silvera and Reynolds [7]. An earlier review by Silvera and Walraven
gives a detailed description of the general development of the field up
to about 1985, with descriptions of the experimental techniques and an
exhaustive treatment of recombination and relaxation processes. The
review by Greytak and Kleppner [8] discusses many of the properties of
the weakly interacting Bose gas in the region of BEC.
where N = 0,1,2, ... , and \N) denotes the state of the magnetic mode
with N photons before coupling to the atom. The angle 6 is given
by tan 20 = cor/S, where cor is the Rabi frequency (cor = yeBrf for a
circularly polarized microwave field of amplitude Brf) and d = co — COQ is
the detuning of the microwave frequency from the resonance frequency,
co0 = Bo/ye (Bo is the applied static field). The energies are
l }
E2N= Nhco - fcO/2(b)
Spin-Polarized Hydrogen: BEC and 2D Superfluidity 165
o.u
Static trap at B=5.0T
5.0
4.0 / \
J 3.0 I :
b
£- 2.0
O
Static trap at B=0.05T •
1.0
r\r\
-4.0 -3.0 -2.0 -1.0 0.0 1.0 2.0 3.0 4.0
8/(0 (dimensionless)
Fig. 2. The spin relaxation loss rate in a microwave trap for hydrogen. The
dashed lines show the loss rate in static traps for two different magnetic field
amplitudes. By using a negative detuning the loss rate can be made negligible in
the microwave trap; for large positive detuning the microwave rate approaches
the static rate.
where Q = (co2 + (52)1/2. The energy levels vary with position through
the variation of Q on position. From Eq. (5) we see that state |2N)
is a trapped state and |1N) is an antitrapped state. By varying S,
the admixture of states can be varied, and, with a reasonable negative
detuning, the character of the state \2N) becomes almost completely \a) 9
the ground state. Agosta et al. [12] have calculated the loss rate due to
relaxation and find that it becomes negligibly small, as shown in Fig. 2.
Thus, with the MWT, the principle loss mechanism of the static trap,
spin relaxation, can be suppressed to a very low level. Moreover, the
relaxation explosion is not a characteristic of this trap. After loading the
trap, by further detuning, the potential walls are lowered and evaporative
cooling can be used to achieve very low temperatures; alternatively atoms
can be selectively pumped from the trapped to the antitrapped state for
forced evaporative cooling, which does not change the shape of the trap.
If BEC can be achieved the principle loss mechanism will be due to
three-body recombination in the center of the trap which will become
important in the density region of 1017 to 10 18 /cm 3 . Since the initial
densities for BEC in this trap will be of order 10 14 /cm 3 , there should
be no problem with a recombination "explosion" when BEC is achieved.
166 /. F. Silvera
MOT/molasses
lasern
(AW input
1 |XW detect
/
\
Cs Oven
Slowing 1
laser • 0 O(
f/cco
Kl ^ Field
—-^ camera
Coil
^ I
Probe
laser
Mirror
Photo
d iode
Fig. 3. A schematic diagram of the cesium trap, showing the microwave cavity
and the several laser beams used for cooling the atoms to be loaded into the
trap. Atoms could be detected by releasing them from the trap and observing
theirfluorescenceas they passed through a probe beam. An alternative means of
observation is to laser illuminate trapped atoms and observe them in fluorescence
with a ccd camera.
Silvera et al. [13] have briefly discussed the problem of atomic losses
due to microwave noise; this appears to be a technically resolvable
problem.
The principle difficulty with the MWT is that it can only be loaded
with very cold atoms. Silvera and Reynolds [14] proposed a hybrid
trap in which H atoms would first be cooled in a static trap (in, say,
the d-state), evaporatively cooled to a few hundred microkelvin, and
then transferred to the MWT. Since this is a very demanding experi-
ment, the MWT was recently demonstrated experimentally [6, 15] with
cesium atoms in an experimental setup shown in Fig. 3. The atoms
are first laser cooled to a few microkelvin using a magneto-optical trap
and polarization laser cooling; the laser is then turned off and the
microwave power, up to 80 Watts in an x-band microwave cavity, is
turned on. A complication in this experiment was that the microwave
potential was insufficiently deep to hold atoms against the gravitational
field. This problem was solved by using a static gradient magnetic field
Spin-Polarized Hydrogen: BEC and 2D Superfluidity 167
To Mixing Chamber
I Cell cooling
Discharge cooling
DISCHARGE
Race-tracks
Al olloy former
Fig. 4. The hybrid static microwave trap under construction at Harvard. The
static trap consists of 11 separately controlled superconducting magnets, used
for trapping, cooling and spatially manipulating the atoms. After being cooled,
atoms will be transferred to the microwave trap by varying the magnetic field
gradients.
The quest for Bose condensation in hydrogen has been focused strongly
on the three-dimensional gas. The attempts to Bose condense hydrogen
as a high density gas, to date, have been thwarted by recombination
heating. Although there are some proposals for experiments which can
handle the heating problem [6] , efforts have moved in the direction of
very low density traps. With regard to two-dimensional KT superflu-
idity, quasi-Bose condensation there seems to have been few efforts in
this direction [17]. Only recently have researchers become aware that it
is probably much easier to produce hydrogen in this state experimen-
tally than three-dimensional BEC. There have been a number of recent
theoretical and experimental results which indicate that this avenue of
research may be the more propitious approach.
The principle of the experiment is to create a two-dimensional sur-
face (of helium on a substrate); hydrogen will condense on the surface
and be stabilized by a magnetic field. The hydrogen behaves as a
two-dimensional gas, populating states acording to the Bose-Einstein
distribution function. However, it has been shown that in two dimen-
sions the Bose-Einstein transition does not take place [18]. Bagnato and
Kleppner [19] studied the ideal Bose gas in two dimensions and found
that, for example, in a non-uniform field with a harmonic potential a BEC
transition does take place. However, Shevchenko [20] studied the same
problem for the weakly interacting gas and found that in this case the
BEC transition does not take place, but, rather, there is a KT transition
and that the critical temperature is very close to the BEC temperature
in the absence of interactions. He then studied the quasi-condensate and
its development in a finite system [21]. The quasi-condensate had been
studied earlier by Kagan et al. [22]. Stoof and Bijlsma [23] have also
studied the thermodynamic properties of the weakly interacting Bose gas
and arrive at similar conclusions. Thus, on theoretical grounds we con-
clude that in two dimensions in an inhomogeneous magnetic field there
is a KT transition and a quasicondensate which becomes macroscopic at
a temperature close to TC2D.
Earlier, it was believed that it would be very difficult to study two-
dimensional BEC because of the very rapid recombination at the surface
coverages which are required and the resultant heating of the sample.
However, recent experiments have shown that when molecules recombine
on a surface there is a high probability that they will form in highly
excited states of Hi and fly off the surface. The fraction F of the
Spin-Polarized Hydrogen: BEC and 2D Superfluidity 169
recombination energy deposited on the surface is found to have an upper
bound of 0.04 for two-body recombination, and there is probably a
similar bound for three-body surface recombination [17, 24]. Thus, only
a very small fraction of the recombination energy goes into the surface
to heat the sample and the heating problem of three-dimensional gases
can be strongly suppressed.
On the theoretical side there are a few additional advantages for the
experimentalist in studying properties of a quantum degenerate system
in two dimensions. Kagan et al. [22] had shown that the recombination
rate constant, K, is reduced by a factor of 6 for three-body recombina-
tion between atoms in the condensate; for a fixed surface coverage the
recombination rate should decrease by this amount as T/T?D decreases
and almost all of the atoms are in the quasi-condensate or superfluid
fraction. The origin for this decrease is in the very simple form of the
wavefunction for the zero-momentum or ground state particles, which
leads to a ^ 3 ! reduction in the matrix element for the process. For
the same reason, the two-particle interaction potential is reduced by 2!
As a result, if there is thermodynamic equilibrium between surface state
atoms and a thermal reservoir of gas atoms in three-dimensions, then at
the transition there is an increase in surface density by about a factor
of two [25]. Since the recombination rate is proportional to KG3, the
rate is expected to have a slight increase at TC2D rather than a decrease.
The most recent development is a calculation by Stoof and Bijlsma [26],
who go beyond the approach of Kagan et al. [22] and, in addition to
wavefunction symmetry effects, they consider the many-body states of
the gas in greater detail. They find an enormous reduction of the rate
constant, more than 400, and an increase in the density step to about
3 at the critical temperature. Together, this predicts that the rate will
decrease by about 33/400 « 15 at the transition. This is likely an over-
estimate since the calculation is probably not valid in the critical region.
Nevertheless, outside the critical region there is still a large predicted
reduction in the recombination rate in the superfluid, according to this
result. From an experimental point of view these predictions suggest
a very nice property of the sample which can be used to identify the
transition, the recombination rate or the recombination heating of the
sample. It will certainly be interesting if such enormous changes (400)
can arise due to Bose effects.
Experimental attempts to study the KT transition are so meager that it
is not meaningful to assess how close experiments have approached this
transition. An experiment was started at the University of Amsterdam to
170 /. F. Silvera
observe the KT transition by optical methods using a Lyman-alpha laser
source. The experimental plan was to introduce a gas of hydrogen to a
small, cold helium surface on the surface of a large reservoir of atoms
stored at a higher temperature, with higher temperature walls. In this way
the main recombination losses would be on the cold surface and could be
maintained at a manageable level. The density of atoms on the surface
would be determined using the Lyman-alpha radiation to optically excite
the system such that the surface coverage could be measured. However, it
was realized that the hydrogen interacts with the surface via the surface
ripplons of the helium and at the low temperature of the experiment
these are weakly coupled to the helium phonon bath [27]. As a result
the ripplons heat up and the hydrogen would not be at a sufficiently low
temperature required for the KT transition; the experiment was aborted.
An experiment at the University of Kyoto [17] using a cold surface has
been carried out at a measured cell temperature of 87 mK; however the
coverages were much too low for the KT transition. This experiment
also determined that the F factor described above is about 0.015. An
experiment is currently being developed at Harvard University [28] using
a two-dimensional magnetic field gradient with a maximum field at the
center of a disc-shaped surface. In this way very high surface coverages
can be achieved in a very small area so that the total sample loss is
very small. The atoms that cover the surface are fed from a large buffer
volume at a higher temperature in such a way that the total number
of atoms on the surface is maintained constant; this inhibits the density
step increase at the critical temperature and should amplify sensitivity
to changes in recombination energy. The means of detection of the
KT transition will be a measurement of the recombination heating on
the cell surface. Similar experiments are planned at the University of
Amsterdam [29] and the University of Turku in collaboration with the
Kurchatov Institute [30], using optical detection or thermal detection
respectively.
4 Conclusions
For more than 13 years, since the initial stabilization of atomic hydro-
gen, researchers have been devising schemes to observe Bose-Einstein
condensation in the weakly interacting Bose gas. In the first several years
a number of clever experiments were attempted, only to be thwarted by
the rapid loss of sample, and heating due to recombination and relax-
ation among hyperfine levels. In this period, there has been an enormous
Spin-Polarized Hydrogen: BEC and 2D Superfluidity 171
growth in the basic theoretical and experimental understanding of this
seemingly simple system. Fortified with this knowledge, new experiments
and approaches have been developed, bringing us within reach of the
goal of BEC, which at times has seemed unattainable. The coming years
promise exciting new results in this simple Boson gas.
References
[I] I.F. Silvera and J.T.M. Walraven, Phys. Rev. Lett. 44, 164 (1980).
[2] I.F. Silvera and J.T.M. Walraven, Progress in Low Temp. Physics,
Vol. X, D.F. Brewer, ed. (Elsevier, Amsterdam, 1986), p. 139.
[3] A. Lagendijk, I.F. Silvera, and B.J. Verhaar, Phys. Rev. B 33, 626
(1986).
[4] See T.J. Greytak, this volume.
[5] D.O. Edwards, Physica B 109 and 110, 1531 (1982).
[6] I.F. Silvera, J. Low Temp. Phys. 89, 287 (1992).
[7] I.F. Silvera and M. Reynolds, J. Low Temp. Phys. 87, 343 (1992).
[8] T.J. Greytak and D. Kleppner, New Trends in Atomic Physics, Vol.
II, G. Grynberg and R. Stora, eds. (North Holland, Amsterdam,
1984), p. 1127.
[9] J.M. Doyle, J.C. Sandberg, LA. Yu, C.L. Cesar, D. Kleppner, and
T.J. Greytak, Phys. Rev. Lett. 67, 603 (1991).
[10] T.W. Hijmans, Y. Kagan, G.V. Shlyapnikov, and J.T.M. Walraven,
preprint (1993).
[II] C.C. Agosta and I.F. Silvera, Spin Polarized Quantum Systems,
S. Stringari, ed. (World Scientific, Singapore, 1989), p. 254.
[12] C.C. Agosta, I.F. Silvera, H.T.C. Stoof, and B.J. Verhaar, Phys. Rev.
Lett. 62, 2361 (1989).
[13] I.F. Silvera, C. Gerz, L.S. Goldner, W.D. Phillips, M.W. Reynolds,
S.L. Rolston, R.J.C. Spreeuw, and C.I. Westbrook, LT20, Physica
B 194^196, in press (1994).
[14] I.F. Silvera and M. Reynolds, Physica B 169, 449 (1991).
[15] L. Goldner, R.J.C. Spreeuw, C. Gerz, W.D. Phillips, M.W. Reynolds,
S.L. Rolston, I.F. Silvera, and C.L Westbrook, LT20, Physica B 194-
196, in press (1994).
[16] Z. Zhao, I.F. Silvera, and M. Reynolds, J. Low Temp. Phys. 89, 703
(1992).
172 /. F. Silvera
[17] A. Matsubara, T. Arai, S. Hotta, J.S. Korhonen, T. Mizusaki, and
A. Hirai, this volume.
[18] RC. Hohenberg, Phys. Rev. 158, 383 (1967).
[19] V. Bagnato and D. Kleppner, Phys. Rev. A 44 , 7439 (1991).
[20] S.I. Shevchenko, Sov. Phys. JETP 73, 1009 (1991).
[21] S.I. Shevchenko, Sov. J. Low Temp. Phys. 18, 223 (1992).
[22] Y. Kagan, B.V. Svistunov, and G.V. Shlyapnikov, Sov. Phys. JETP
66, 314 (1987).
[23] H.T.C. Stoof and M. Bijlsma, Phys. Rev. B 47 (2), 939 (1993).
[24] Z. Zhao, E.S. Meyer, B. Freedman, J. Kim, J.C. Mester, and I.F. Sil-
vera, LT20, Physica B 194-196, in press (1994).
[25] B.V. Svistunov, T.W. Hijmans, B.V. Shlyapnikov, and J.T.M. Wal-
raven, this volume.
[26] H. Stoof and M. Bijlsmaa, preprint (1993).
[27] M.W. Reynolds, I.D. Setija, and G.V. Shlyapnikov, Phys. Rev. B 46,
575 (1992).
[28] I.F. Silvera, unpublished.
[29] M. Reynolds, private communication.
[30] S. Jaakkola, private communication.
Laser Cooling and Trapping of
Neutral Atoms
Y. Castin, J. Dalibard and C. Cohen-Tannoudji
Laboratoire de Spectroscopie Hertzienne de VE.N.S.'f
et College de France
24, rue Lhomond
F-75231 Paris Cedex 5
France
Abstract
We present a simple review of the basic physical processes allowing one to
control, with laser beams, the velocity and the position of neutral atoms.
The control of the velocity corresponds to a cooling of atoms, that is, a re-
duction of the atomic velocity spread around a given value. The control of
the position means a trapping of atoms in real space. The best present per-
formances will be given, in terms of the lowest temperatures and the highest
densities. The corresponding highest quantum degeneracy will also be esti-
mated. It is imposed by fundamental limits, which will be described briefly .
We also give the general trends in this field of research and outline the new
directions which look promising for observing quantum statistical effects in
laser cooled atomic samples, but which are for the moment restricted by
unsolved problems.
173
174 Y. Castin, J. Dalibard and C. Cohen-Tannoudji
1.1 Effect of Light on the Atomic Internal State
Since the atoms in laser cooling are illuminated by quasi-resonant light,
they can be considered as two-level atoms, with a metastable state
g, called "ground state" in what follows, and an excited state e. The
transition between g and e has a resonant frequency COA and an electric
dipole d. The excited state can decay by spontaneous emission with rate
F. In this section, we consider the simple case of atoms with no sublevels
in the ground state g. The influence of the presence of Zeeman sublevels
in g on laser cooling is the subject of Section 2.
The laser electric field is a superposition of travelling waves of fre-
quency coL close to COA and of complex amplitude SQ. An important
parameter is 5 = COL — COA, the detuning of the laser frequency CDL from
the atomic frequency COA. The dipolar coupling between atom and light
leads to a time dependent amplitude of transition from g to e, denoted by
(Q/2)e-i0JL\ and from e to g, (Q*/2) eia)Lt. The parameter Q = -dS0/h is
called the Rabi frequency. The amplitude of transition can be made time
independent by the following time dependent unitary transformation S(t)
changing the excited state:
which gives the |g) — • \e) transition the effective Bohr frequency COA —
CDL = — S instead of <X>A> This leads to the time independent effective
hamiltonian ifeff for the evolution of the internal atomic state:
-d-iY/2 0/2 \
(2)
"72 0 )•
This matrix is given in the {|e),|g)} basis and the — iT/2 term accounts
for the instability of the excited state.
The effect of the coupling will be considered in the perturbative regime:
Fz = -<xvZ9 (4)
k,
vww* CJ T
WWV* O *AAAAA
cj L -k L v co L +k L v
Fig. 1. The laser waves of the Doppler cooling configuration, in the laboratory
(a) and in the atomic rest frame (b).
kBTD = —. (8)
0
/ A/8 A/4 3A/8 A/2 z
y/
Fig. 3. Resulting electric field in the x-y laser configuration.
led to a new picture for the laser cooled atomic samples, the so-called
"optical lattices".
J=3/2
(a)
Fig. 4. (a) Atomic-level scheme and intensity factors (square of the Clebsch-
Gordan coefficients) for a jg — 1/2 —• je = 3/2 atomic transition, (b) In the
Sisyphus cooling configuration, the position-dependent light shifts of the ground
state sublevels and a typical atomic trajectory in the position-energy space.
state and an angular momentum je = 3/2 in the excited state. This atom
is illuminated by the previous field configuration, with a laser frequency
a>L below the resonance frequency COA :
Uo = -^hds (14)
1000
The scaling law kBT ~ Uo is recovered in the limit of large U0/ER. The
existence of a threshold on Uo corresponds to a rapid increase of the
average kinetic energy, when UO/ER becomes too small. The smallest
possible value of the root mean square atomic momentum is of the order
of 5.5ft/c,reached for Uo ~ 100ER in the limit of large detunings \S\ > T.
Fig. 6. Band structure of the energy spectrum for the quantum motion in U+(z)
optical potential wells. The vibrational levels are labelled v = 0,1,.... The dark
circular areas are proportional to the steady state populations of the vibrational
levels.
free motion well above the potential hills. The effect of optical pumping
in this basis can be described simply by transition rates between the
various energy levels, when the inequality (16) holds. The corresponding
rate equations allow one to calculate the steady state populations of the
energy bands. The maximal population in the most populated band v = 0
is found to be 0.34 in [21], for Uo ^ 60ER.
The existence of well resolved external quantum levels in ID Sisyphus
cooling has been demonstrated experimentally. A probe laser beam, with
a very small intensity, is sent through the optical molasses along the
direction of the laser cooling beams. Two side-bands are clearly observed
on the probe absorption spectrum, for a probe frequency cop = COL ±^OSC-
They correspond to stimulated Raman transitions between two successive
vibrational levels in the potential wells. Since the lowest level is the most
populated one, Raman cycles starting with atoms in the state v = 0
are expected to be dominant. When cop is below coL (cop ~ coL — £losc\
the dominant process is therefore the absorption of one photon in the
cooling beams followed by stimulated emission of one photon in the
probe beam, and the probe beam is amplified. When cop is above COL
(cop ~ coL + Qosc), the Raman processes are mainly the absorption of
one photon in the probe beam followed by the stimulated emission of
one photon in the cooling beams, and the probe beam is absorbed (see
Fig. 7 taken from [22]). The so-called "overtones" (Raman transitions
186 Y. Castin, J. Dalibard and C. Cohen-Tannoudji
n^ = o ^(njl£ o ) 2 , (17)
The geometry of the magneto-optical trap has been described via a simple
ID model in Section 1.4. It can be extended to a 3D configuration in the
following way. The magnetic field gradient is provided by two coaxial
coils travelled by opposite electric currents. Close to r = 0, where the
resulting magnetic field is vanishing, one has the approximate spatial
dependence for B:
(18)
<, t ~ 1. (20)
190 Y. Castin, J. Dalibard and C. Cohen-Tannoudji
This corresponds to n ~ 2 • 1012 atoms/cm 3 on cesium, of the order of
the experimental results.
Fig. 8. The A atomic system in the 0+—c_ laser configuration. In the absence
of spontaneous emission, note the existence of closed families of atomic states
labelled by p, the atomic momentum along z in the excited state.
the atomic motion along z by introducing the external states \p) corre-
sponding to the plane waves of momentum pez. The atomic dynamics
can then be analyzed as follows [40].
In the absence of spontaneous emission, an atom initially in the state
\eo,p), i.e. in the internal (electronic) state |eo) and with a center of
mass motion of momentum p along z, is coupled only to |g_,p — hk)
(by stimulated emission of a photon in the a+ wave) with an amplitude
a+, and to \g+,p + hk) (by stimulated emission of a photon in the
<j_ wave) with an amplitude a_. The space of atomic states can then
be split into a direct sum of an infinite number of closed families
^(p) = {\eo,p),\g->P — bk)>\g+>P + hk)} labelled by p. Since there is
only one internal excited state \eo) for two ground levels |g+), we can
form, inside each family ^(p), a particular linear combination \ipwc(p))
uncoupled to the laser:
[ l J f c ) | + ftfc)] (21)
N(p)
pulses and provide a high selectivity in atomic velocity. The atoms whose
velocity v meets the Raman resonance condition v = vo are transferred
into 12) by the pulses and are then "pushed" towards zero velocity back
into |1) through a pumping resonant traveling wave. Successive cooling
cycles of this type are then performed, with a scanning of vo from high
velocities towards low velocities, in order to collect most of the atoms.
Since vo is always different from 0, there is a small region around v = 0
where the atoms never meet the Raman resonance condition, and where
they pile up. Very sharp momentum distributions around p = 0 have thus
been obtained and observed in ID, with effective temperatures down to
TR/10.
4 What Next?
As explained in Section 3, the subrecoil cooling schemes known at present
cannot be applied efficiently in 3D because of gravity. In this section, we
explain how to make an atomic cavity by using gravity and a parabolic
atomic mirror (Section 4.1). A quantum picture for the atomic motion
inside this cavity is briefly given in Section 4.2: as light waves do in usual
laser cavities, matter waves have well defined modes in the gravitational
cavity, and one may hope to obtain quantum statistical effects if several
atoms occupy the same quantum modes.
indeed have a rate scaling as I/(COL — &>A)2- However, COL has to be not
too far from resonance so that the laser has a sufficiently high intensity to
provide reactive effects (i.e. light shifts, proportional to //(COL — &>A)) able
to reflect the atoms. This last condition is very restrictive for thermal
atomic velocities, and in this case only reflection for grazing atomic
incidence is observed [50]. On the contrary, it is easily fulfilled using
laser cooling techniques, and reflection under normal atomic incidence
can then be obtained [51].
In order to bound the transverse atomic motion also, the plane mir-
ror of the previous scheme is replaced by a parabolic mirror, with an
evanescent wave still present at its surface. The atoms are released with
negligible initial velocities from a switched-off magneto-optical trap and
bounce off the evanescent wave (see Fig. 10). The corresponding clas-
sical orbits are paraxial and are shown to be stable [52]. Experimental
realizations are in progress in Paris (at the ENS) and in Gaithersburg
(at NIST). Recent observation of about ten bounces has been reported
in Paris for cesium atoms with the following experimental evidence. At
time t after the drop of the atoms, a probe beam is sent through a
small volume d V around the initial position of the atomic cloud, and the
fluorescence light is collected. This procedure gives a signal proportional
to the number N(t) of atoms in the volume dV. Such a measurement
is destructive because atoms are pushed away and heated by the probe
196 Y. Castin, J. Dalibard and C. Cohen-Tannoudji
200
Time in ms
Fig. 11. The number of atoms in the gravitational cavity in Paris is measured
as a function of time using thefluorescenceinduced by a probe beam (see text).
The points on the diagram give the number of atoms in the probe beam for
different times after their release. The curve is a fit calculated by a Monte-Carlo
simulation. The successive bounces are labelled 1,2... .
beam so that the whole process has to be restarted from the beginning if
one wishes to obtain N(tf) at a different time t''. The result of numerous
such cycles is shown in Fig. 11 (see [53]; less recent results, with only
four bounces, are given in [54]).
Fig. 12. The calculated probability distribution in real space of a mode in the
gravitational atomic cavity. The cross corresponds to the focus of the parabolic
surface of the mirror.
References
[1] J. Dalibard and C. Cohen-Tannoudji, J. Opt. Soc. Am. B 2, 1707
(1985).
[2] S. Chu, J. Bjorkholm, A. Ashkin and A. Cable, Phys. Rev. Lett. 57,
314 (1986).
[3] P. Gould, P. Lett, P. Julienne, W.D. Phillips, W. Thorsheim and J.
Weiner, Phys. Rev. Lett. 60, 788 (1988).
[4] T. Hansch and A. Schawlow, Opt. Commun. 13, 68 (1975).
[5] D. Wineland and H. Dehmelt, Bull. Am. Phys. Soc. 20, 637 (1975).
[6] S. Chu, L. Hollberg, J. Bjorkholm, A. Cable and A. Ashkin, Phys.
Rev. Lett. 55, 48 (1985).
[7] D J . Wineland and W.M. Itano, Phys. Rev. A 20, 1521 (1979).
[8] J.P. Gordon and A. Ashkin, Phys. Rev. A 21, 1606 (1980).
Laser Cooling and Trapping of Neutral Atoms 199
[9] E.L. Raab, M. Prentiss, A. Cable, S. Chu and D.E. Pritchard, Phys.
Rev. Lett. 59, 2631 (1987).
[10] P. Lett, R. Watts, C. Westbrook, W.D. Phillips, P. Gould and H.
Metcalf, Phys. Rev. Lett. 61, 169 (1988).
[11] P. Lett, W.D. Phillips, S. Rolston, C. Tanner, R. Watts and C.
Westbrook, J. Opt. Soc. Am. B 6, 2084 (1989).
[12] J. Dalibard, C. Salomon, A. Aspect, E. Arimondo, R. Kaiser,
N. Vansteenkiste and C. Cohen-Tannoudji, Atomic Physics 11, S
Haroche, J.C Gay and G. Grynberg, eds. (World Scientific, Singa-
pore, 1989), p.199.
[13] Y. Shevy, D.S. Weiss, P.J. Ungar and S. Chu, Phys. Rev. Lett. 62,
1118 (1989).
[14] D.S. Weiss, E. Riis, Y. Shevy, P.J. Ungar and S. Chu, J. Opt. Soc.
Am. B 6, 2072 (1989).
[15] C. Salomon, J. Dalibard, W.D. Phillips, A. Clairon and S. Guellati,
Europhys. Lett. 12, 683 (1990).
[16] C. Gerz, T.W. Hodapp, P. Jessen, K.M. Jones, W.D. Phillips, C.J.
Westbrook and K. M0lmer, Europhys. Lett. 21, 661 (1993).
[17] J. Dalibard and C. Cohen-Tannoudji, J. Opt. Soc. Am. B 6, 2023
(1989).
[18] PJ. Ungar, D.S. Weiss, E. Riis and S. Chu, J. Opt. Soc. Am. B 6,
2058 (1989).
[19] C. Cohen-Tannoudji, in Fundamental systems in Quantum optics,
Proceedings of Session LIII of Les Houches Summer School, J.
Dalibard, J.-M. Raimond and J. Zinn-Justin, eds. (North-Holland,
Amsterdam, 1992).
[20] Y. Castin, J. Dalibard and C. Cohen-Tannoudji, Light Induced Ki-
netic Effects, L. Moi, S. Gozzini, C. Gabbanini, E. Arimondo and
F. Strumia, eds. (ETS Editrice, Pisa, 1991).
[21] Y. Castin and J. Dalibard, Europhys. Lett. 14, 761 (1991).
[22] J.Y. Courtois, these de doctorat de l'Ecole Polytechnique, Paris,
France, 1993.
[23] P. Verkerk, B. Lounis, C. Salomon, C. Cohen-Tannoudji, J.Y. Cour-
tois and G. Grynberg, Phys. Rev. Lett. 68, 3861 (1992).
[24] J.Y. Courtois and G. Grynberg, Phys. Rev. A 46, 7060 (1992).
[25] PS. Jessen, C. Gerz, P.D. Lett, W.D. Phillips, S.L. Rolston, R.J.C.
Spreeuw and C.I. Westbrook, Phys. Rev. Lett. 69, 49 (1992).
[26] K. Berg-Sorensen, Y. Castin, K. M0lmer and J. Dalibard, Europhys.
Lett. 22, 663 (1993).
200 Y. Castin, J. Dalibard and C. Cohen-Tannoudji
[27] Y. Castin, K. Berg-Sorensen, K. Molmer and J. Dalibard, to appear
in Fundamentals of Quantum Optics HI, Ehlotzky, ed. (Springer,
Berlin, 1993).
[28] A. Hemmerich, Zimmerman and T.W. Hansch, Europhys. Lett. 22,
89 (1993).
[29] A. Hemmerich and T.W. Hansch, Phys. Rev. Lett. 70, 410 (1993).
[30] G. Grynberg, B. Lounis, P. Verkerk, J.Y. Courtois and C. Salomon,
Phys. Rev. Lett. 70, 2249 (1993).
[31] A. Clairon, P. Laurent, A. Nadir, M. Drewsen, D. Grison, B. Lounis
and C. Salomon, in the Proceedings of the 6th European Frequency
and Time Forum, held at ESTEC, Noordwijk (ESA SP-340, June
1992).
[32] A.M. Steane and C.J. Foot, Europhys. Lett. 14, 231 (1991).
[33] A.M. Steane, M. Chowdhury and C.J. Foot, J. Opt. Soc. Am. B 9,
2142 (1992).
[34] C D . Wallace, T.P. Dinneen, K.Y.N. Tan, A. Kumarakrishnan, PL.
Gould and J. Javanainen, to be published (1993).
[35] T. Walker, D. Sesko and C. Wieman, Phys. Rev. Lett. 64,408 (1990).
[36] A.M. Smith and K. Burnett, J. Opt. Soc. Am. B 9, 1240 (1992).
[37] A. Gallagher and D.E. Pritchard, Phys. Rev. Lett. 63, 957 (1989).
[38] PS. Julienne and J. Vigue, Phys. Rev. A 44, 4464 (1991).
[39] A. Aspect, E. Arimondo, R. Kaiser, N. Vansteenkiste and C. Cohen-
Tannoudji, Phys. Rev. Lett. 61, 826 (1988).
[40] A. Aspect, E. Arimondo, R. Kaiser, N. Vansteenkiste and C. Cohen-
Tannoudji, J. Opt. Soc. Am. B 6, 2112 (1989).
[41] M. Kasevich and S. Chu, Phys. Rev. Lett. 69, 1741 (1992).
[42] H. Wallis and W. Ertmer, J. Opt. Soc. Am. B 6, 2211 (1989).
[43] K. Molmer, Phys. Rev. Lett. 66, 2301 (1991).
[44] F. Bardou, J.P Bouchaud, O. Emile, A. Aspect and C. Cohen-
Tannoudji, Phys. Rev. Lett. 72, 203 (1994).
[45] F. Mauri, F. Papoff and E. Arimondo, in Light Induced Kinetic
Effects, L. Moi, S. Gozzini, C. Gabbanini, E. Arimondo and F.
Strumia, eds. (ETS Editrice, Pisa, 1991).
[46] M.A. Ol'shanii and V.G. Minogin, in Light Induced Kinetic Effects,
L. Moi, S. Gozzini, C. Gabbanini, E. Arimondo and F. Strumia, eds.
(ETS Editrice, Pisa, 1991).
[47] J.P. Bouchaud and A. Georges, Phys. Rep. 195, 125 (1990).
[48] B. Lounis, J. Reichel and C. Salomon, C. R. Acad. Sci. Paris, Serie
II, 316, 739 (1993).
[49] R. Cook and R. Hill, Opt. Comm. 43, 258 (1982).
Laser Cooling and Trapping of Neutral Atoms 201
[50] V. Balykin, V. Letokhov, Y. Ovchinikov and A. Sidorov, Phys. Rev.
Lett. 60, 2137 (1988).
[51] M. Kasevich, D. Weiss and S. Chu, Opt. Lett. 15, 607 (1990).
[52] H. Wallis, J. Dalibard and C. Cohen-Tannoudji, Appl. Phys. B 54,
407 (1992).
[53] C. Aminoff, A. Steane, P. Bouyer, P. Desbiolles, J. Dalibard and C.
Cohen-Tannoudji, to be published.
[54] C. Aminoff, P. Bouyer and P. Desbiolles, C. R. Acad. Sci. Paris
Serie II, 316, 1535 (1993).
10
Kinetics of Bose-Einstein Condensate
Formation in an Interacting Bose Gas
Yu. Kagan
Department of Superconductivity and Solid State Physics
Kurchatov Institute
123182 Moscow
Russia
Abstract
The kinetics of the formation of coherent correlation properties associated
with Bose condensation is studied in detail. The evolution of a nonequilib-
rium state with no "condensate-seed" is related to a hierarchy of relaxation
times. At the first stage, a particle flux in energy space toward low energies
sets in. The evolution in this case is described by a nonlinear Boltzmann
equation, with a characteristic time given by interparticle collisions. When
the particles which will later form the condensate have a kinetic energy
which is less than the potential energy, a quasicondensate starts to form.
In this stage, fluctuations of the density (but not of the phase) are sup-
pressed and short-range coherent correlation properties are governed by
the equation of motion for a quasiclassical complex field. The next stage is
connected with the formation of the long-range order. The time for forming
topological order and therefore genuine superfluidity proves to be dependent
on the system size. The off-diagonal long-range order, arising after the at-
tenuation of long-wave phase fluctuations, has a size-dependent relaxation
time as well.
1 Introduction
The problem of Bose-Einstein Condensation (BEC) kinetics, being in-
teresting in itself, has acquired a special significance in connection with
experimental efforts to observe this condensation in a number of systems
with particles with a finite lifetime. Such systems include spin-polarized
atomic hydrogen [1], excitons [2] and biexcitons [3] in semiconductors
and, more recently, laser-cooled atomic systems [4]. In all cases, the
density is supposed to be low so that we always deal with a weakly inter-
202
Kinetics of BEC Formation in an Interacting Bose Gas 203
acting Bose gas. This makes the investigation of these systems especially
attractive, because it allows the possibility of revealing the main features
which are characteristic for a Bose system in general, and, at the same
time, provides a chance for direct comparison with theory.
In all conceivable metastable systems whose evolution should lead
to BEC, the initial state contains no seed of condensate. If the final
equilibrium state is one with a sufficiently high condensate density, then
we are faced with the time evolution of an essentially nonequilibrium
system. Under these conditions the question of the time for attaining BEC,
or, more precisely, of the formation time of the appropriate correlation
properties, becomes very nontrivial. The answer to this question is
important not only for comparison with the lifetime of a system, but it
can also be crucial for the possibility of detecting the condensate.
In Ref. [5] it has been shown that the rate for inelastic processes
decreases substantially when a Bose condensate appears in the system.
This effect opens up an interesting opportunity for the observation of
BEC, particularly in an atomic hydrogen gas. Although there is no
true Bose condensate for T ^= 0 in the two-dimensional case, it was
shown in Ref. [6] that the change in the probability for the inelastic
processes persists below the Kosterlitz-Thouless transition, due to specific
properties of the two-dimensional quasicondensate (i.e. a condensate with
a fluctuating phase). However, in all cases this effect manifests itself only
after the development of specific quantum correlations in the process of
evolution of the system.
In a nonuniform space the opposite effect is realized: the appearance of
a condensate causes a sharp increase of the rate of any density-dependent
process (see, e.g. Ref. [7]). It allows BEC to be revealed even when the
condensate is localized in a very small volume. The latter is characteristic,
in particular, of a trap-type experiment (see, e.g. Ref. [8]). Again, we
need a certain time for the necessary quantum correlations to form.
One promising way of studying BEC in an excitonic gas is based on
the measurement of the transport motion of an expanding gas with the
aim of revealing the excitonic superfluidity [2,9]. The use of this method
assumes indirectly that the time of the formation of topological order is
short enough.
Several theoretical papers have addressed the question of the time evo-
lution of a nonequilibrium interacting Bose gas. Levich and Yakhot [10],
considering the evolution of an ideal Bose gas coupled to a heat bath,
came to the conclusion that the formation of a condensate peak requires
an infinite time. Nozieres [11] also obtained the same result. Recently,
204 Yu. Kagan
studying the kinetics of BEC in an ideal Bose gas, Tikhodeev [12] has
obtained the same answer, provided the size of the system goes to infinity.
Later, Levich and Yakhot [13] took into consideration the interparticle
interactions which should naturally play a dominant role in the temporal
evolution. Within the framework of the nonlinear Boltzmann kinetic
equation, they found a finite time for BEC formation. However, these
authors made a drastic modification of the collision integral in order
to simplify the calculation, which led to such a strong change of the
character of the evolution of the particle energy distribution that the
authors themselves doubted the validity of the results. Moreover, they
realized correctly that the kinetic equation is not applicable in the low
energy region where the kinetic energy becomes less than the potential
energy. The attempt to improve the kinetic equation was not adequate
because it still made use of the random-phase approximation, which fails
precisely in this region.
Snoke and Wolfe [14] undertook a direct numerical evaluation of the
Boltzmann kinetic equation for the weakly interacting Bose gas. The
calculations demonstrated the formation of the Bose distribution with
chemical potential \i close to zero on the time scale of the classical
scattering time of the particles. However, a condensate did not appear
even after a much longer period of time. Simultaneously, the conservation
of the particle number was violated. As we shall see below, both results
are not accidental and have a common origin.
Another extreme result was recently reported by Stoof [15], who
asserted that the time required for the formation of a Bose condensate
was ~ h/TC9 where Tc is the temperature of BEC. That result seems
surprising since it does not depend on the interparticle interaction, and
moreover this time is shorter than any characteristic interaction time in
a dilute Bose gas.
Actually, the question of the formation time of a Bose condensate does
not have an unambiguous answer. The real answer depends strongly on
the particular problem under consideration (see Ref. [16]). The ambiguity
arises at the stage of evolution when a substantial fraction of the particles
which should form the condensate in equilibrium is concentrated in the
energy interval e < noU, where U is the effective particle interaction
vertex, and no is the equilibrium density of the condensate. At this
stage the kinetic energy of the particles becomes smaller than their
average potential energy and a pronounced mixing of states with different
momenta k arises. The state of the system, taking into account that in
this case the occupation numbers obey the condition n^ > 1, can be
Kinetics of BEC Formation in an Interacting Bose Gas 205
characterized by a quasiclassical complex-valued field ¥ (see e.g. [17])
whose dynamical evolution is described by the nonlinear Schrodinger-
type equation (h = 1)
where m is the mass of the particle. At this stage the random phase
approximation is not valid and (1.1) cannot be reduced to an equation for
occupation numbers. (There is a well-known analogy: the classical electric
field and the Maxwell equations as a limit of quantum electrodynamics
when the photon occupation numbers are much larger than unity.)
In the limit t —• oo, the solution of (1.1) approaches the value of the
complex order parameter with a macroscopically defined modulus and
phase which characterize the equilibrium state of an interacting Bose
gas at T < Tc. In the process of evolution, the modulus and the phase
of the function ¥ fluctuate strongly. Only with the relaxation of these
fluctuations do the proper coherent correlations start to set in. The
attenuation of the modulus or the density fluctuations begins first. This
leads to the formation of a short-range order with correlation properties
that are close to those of an equilibrium system with a true condensate
(see below). The characteristic time scale for this stage of the evolution
is given by
and, correspondingly, the length scale of the region where this short-range
order comes into play has the value of the correlation length
(L3)
(!
i%
ot
= -^f
2m
+ noU(\f\2f-f).
In terms of dimensionless variables
icokTc(5nk/n0) = -2r2k2Q>k,
and we have
[Snk/n0]2 = 2r2k2 | % | 2 .
Hence we obtain
E * (rco/m) /
J
k
we find from (2.9)
z = < £ + (r, tyv+(r9 t)* + (r, tyv(T, tyv(r9 tyv(r91) >, (2.11)
as was shown in Ref. [5] This is a typical short-range correlation function.
The coincidence of arguments reflects only the fact that the interaction
takes place at distances much less than rc and times much less than xc.
In the case at hand, this correlation function reduces to
Z=nl< f 3(r, t)/ 3 (r, t) > = < n3 > . (2.12)
If fluctuations in / are suppressed, then
Z=n\. (2.13)
If fluctuations are present, then
Z = n\ + 3n0 < (cSn0)2) > + < (<5n0)3) > . (2.14)
Z = 6n30 . (2.18)
in accordance with the Ginsburg criterion (see, e.g. Ref. [24]). Here n is
the total density of particles, and a is the s-wave scattering length. The
212 Yu. Kagan
latter is connected to U via the well-known relation
The small value of the gas parameter na3 allows (3.1) to be compatible
with the conditions
no/n < 1 (3.3)
and
T —T
-V-•* C
< 1, (3-4)
where T is the equilibrium temperature of the gas. We take advantage
of this possibility to make the discussion more transparent.
For the energy e* which defines the boundary of the nonlinear region,
we have in this case
e* < T. (3.5)
This means that in the first stage the excess particle flux in the energy
space covers a rather wide interval (T,e*), where the deviation of the
distribution function from its equilibrium value is relatively small.
The general kinetic equation for a Bose gas in a spatially uniform
case, for a distribution which is isotropic in terms of momentum, has the
well-known form (nt = ne.)
n2)n3n4l (3.6)
where
W = fr2m3/47r3fc3 = ^ (3.7)
and
. (3.9)
Kinetics of BEC Formation in an Interacting Bose Gas 213
The temporal evolution in this region can be found by means of the solu-
tion of the linearized equation (3.6). The following simple considerations
enable one to reveal the main features of this stage.
We assume that a local quasiequilibrium is reached in the energy
interval where the bulk of excess particles is concentrated. Then the
particle distribution can be written as
(here we have made use of the circumstance that T is close to Tc, and
we have used the relationship between Tc and n). In this case we have
The parameter fi(t) lags behind 6o(O at all times and is comparable to this
quantity when the boundary of the linear regime, 6*, is reached. This cir-
cumstance demonstrates that it is legitimate to assume a quasiequilibrium
distribution of the form given in (3.10) and (3.11).
Let us find an expression for the particle flux in energy space. This
214 Yu. Kagan
flux is related to the motion of the front eo(t). Using (3.11) and (3.13) we
find
Q = -a4 /2 6 O n eo * -ino(W^o). (3.16)
The ratio eo/^o can be found from the kinetic equation (3.6). Analysis
of the collision integral shows that the main contribution is provided
by processes in which all particles have an energy of the same order of
magnitude. Besides, after linearization, the collision-integral operator is a
homogeneous functional of the energy to zeroth order. In other words, it
does not change under the replacement e,- —• >ie,. As a result, the collision
integral is characterized by an energy-independent relaxation time and
we have approximately
£oM> « -7 Akin • (3.17)
Explicit calculations using the kinetic equation (3.6) result in the es-
timate y « 4 (ikin is equal to (1.6) with o = 8na2). The particle flux in
energy space thus takes the simple form
2«yn o/2Tkin. (3.18)
It is interesting to note that the flux has a wave-shape with an increas-
ing front amplitude and a decreasing particle-distribution width during
the passage of the considered energy interval (see (3.11)). Using (3.17) and
(3.14), we can estimate the time which takes the front of the distribution
to reach the boundary of the linear region e* as
( (3.19)
y \nj
The time required to cross the linear region is thus determined by the
ordinary collision time scale Tkin? enhanced by a factor \n(n/no) at small
wo-
**~Wn¥. (4.3)
Analyzing a nonlinear kinetic equation of such a type, Zakharov [22]
found that the equation Jco\\ = 0 has two extra solutions, corresponding
to a steady state particle flux and to a steady state energy flux in the
energy space. In first case, the solution has the form
ne = A/e1/6 . (4.4)
T^(e)~WA2/eV\ (4.5)
Thus, the solution (4.4) leads to the enhancement of the collision processes
with decreasing energy in both the incoming and the outgoing terms.
Considering the particle flux moving toward 6 = 0, there are good rea-
sons to suppose that behind the front the particle distribution should be
close to (4.4). Direct numerical calculations performed by Svistunov [23]
have corroborated this conjecture.
Thus, the effective transformation from the distribution in (3.11) to
that in (4.4) starts at the energy e ~ e*. Once the distribution (4.4)
has been established behind the front ?o(O of a particle-flux wave, the
parameter A can be found from the condition
Note that the restructuring of the distribution from (3.11) to (4.4) leaves a
substantial fraction of the particles in the tail of the distribution (e ~ e*)
Making use of (4.6), one can rewrite (4.5) in the form
3
- (4-7)
216 Yu. Kagan
Writing the kinetic equation as h€ = J+ — J_, we have J+ ~ n€/x^{e\
while for the left-hand side we have h€ ~ ne/Tkin. It means that in the
first approximation Jcoii « 0 at e < e* and, consequently, the solution
(4.4) is self-consistent.
The value of the flux associated with the motion of the front is
determined by a relation such as (3.16). Taking into account (4.4), we
obtain
( 4 .8)
(to is reckoned from the beginning of the nonlinear regime). The quantity
to determines the time at which the front of the particle-flux wave arrives
in the coherent region. Using the value of the flux Q entering from the
linear region (see (3.18)), we immediately conclude to ~ ikin-
In the initial stage, the buildup of particles in the coherent region is
described by
nc*Q(t-t0). (4.11)
he = ncJ'U[n\), (4.13)
where
W 1 f€l
J'con([n],ei) = r ^ { / de2[n1(n2 + ni_ 2 ) - n2n1-2]
a
6j JnoU
+2 de2[nin2-i-n2(ni+n2-i)]}, (4.14)
Jei
Kinetics of BEC Formation in an Interacting Bose Gas 217
The first term on the right-hand side of (4.12) describes the interaction
of particles belonging to the kinetic region before and after a collision.
The distribution (4.4) causes the divergence of the integral in (4.15) at
low energies. If one defines the lower bound of the energy as eC9 one
obtains for the collision integral
= ±2n
Here c is the sound velocity (2.8). Let us first consider the relaxation of
topological defects. The main channel of the relaxation is self-annihilation
of the vortex rings caused by interaction with the elementary excitations
which provide the drag force and thus the dissipation of the energy.
The small-radius rings naturally disappear first. We define the minimum
radius in the distribution (5.1) at a moment t as R(t). Then it is easy to
Kinetics of BEC Formation in an Interacting Bose Gas 219
make sure that the average radius of the rings is equal to R(t) and the
total length L of the vortex line in a unity volume equals
L(t) « J - . (5.5)
R(t)
This means that the relaxation at any moment t is related to the rings of
one size scale R(t).
In reality, the emerging vortex structure may resemble a tangle rather
than an ensemble of the rings. Usually, when studying the kinetics of
such a tangle, the average distance between the vortex lines is considered
as a basic parameter of a system. The same scale defines the characteristic
radius of the curvature. Since this parameter and L are linked by the
relation (5.5), the kinetics of the two models prove to be equivalent. For
simplicity we shall only refer to the first one.
The characteristic time of the self-annihilation of a ring of radius R is
(see, e.g. Ref. [19])
2nps R2
^R)™—]^RF C- (5 6)
"
Here y is the drag force coefficient per unit length of a vortex line. We
have taken into account that the core radius is of the order of rc and
R(t) > rc. It can be shown [18] that in an interacting Bose gas in the
temperature interval
n0U < T < TC9 (5.7)
the coefficient y is equal to
y = p s r o a, (5.8)
where
. (5.11)
Note that this expression differs from (5.11) only in a substantial decrease
of the dependence upon the small interaction parameter a. (In the region
T <n0U one finds that a ~ T5 [27].)
Thus the time of the formation of topological long-range order, which
is determined by (5.12) (or (5.11)), turns out to be dependent on the size
of the system. In a time of the order of TV(DQ), the Bose gas becomes
truly superfluid.
where pn is the density of normal liquid, and s is the entropy per unit
mass. Considering p and T as independent thermodynamic variables, we
can write for a small variation of parameters in a sound wave
Introducing (6.3) into (6.2), we find the system of equations for pr and
V . The equations (6.2)-(6.3) are correct in the general case. Applying
these equations to the interacting Bose gas, we find that the results differ
noticeably from the usual results for superfluid helium.
In the temperature interval (5.7) the normal excitations practically
have the dispersion law of free particles. Consequently, the normal
component has the properties of an ideal Bose gas. In this case, for the
temperature-dependent part of the pressure p T , we have
*0, (6.4)
V UP J T
and as a result
= c\ (6.5)
Taking into account the fact that ps =£ f(p) we find, with allowance for
(6.6), that
3 V
(68)
Using (6.7) and (6.8), we can find the sound dissipation by employing
the standard approach (see, e.g. [28]). Since n o & < 71 and the specific
heat has temperature dependence given by (6.6), it is easy to show that
the role of the viscosity is negligible. As a result, the attenuation of
sound is determined by the thermal conductivity. Direct evaluation gives
for the phonon relaxation time
(6.10)
•t ph(k=TCl) (na3)l'2Tc In R/rc'
For a realistic value of the system size, this ratio is always much greater
than unity. This assertion has an additional basis. As is well known,
sound reflection from a wall is accompanied by strong absorption (see,
e.g. Ref. [28]). At the same time, the phonon free path length approaches
the value Do much earlier than R(t) approaches this size.
Since the phase relaxation time T^ actually coincides with TPJ,, we can
conclude that %$ < xv. This signifies that the phase relaxation happens
faster than the vortex annihilation. Therefore, for the conditions un-
der consideration, complete long-range order should become established
when the topological order appears.
In a system of limited size, if the inequality / > Do is justified, the
hydrodynamic condition (6.1) is violated. In this case, the sound attenu-
ation is determined by the direct scattering of phonons with the normal
Kinetics of BEC Formation in an Interacting Bose Gas 223
fluid excitations. The calculation of this process leads, in the temperature
interval (5.7), to the relation [18]
kT (6.11)
(6 13)
< ~ Wo -
(we have again restored h in the explicit form).
7 Final Comments
In a number of experimental situations, nonuniform conditions are in-
volved and the condensate should form in a small volume. This takes
place, for example, in different kinds of magnetic traps or during the
creation of excitons by a laser pulse. We make some remarks in this
connection.
The formation of the quasicondensate in this case does not differ
noticeably from the uniform case characterized by the same scale of
relaxation time. The number of vortices is now limited, especially at low
densities. The time of their self-annihilation drastically shortens due to
the small size of the condensate location region. The free expansion of
the system, which is characteristic in particular of the excitonic gas in a
semiconductor, causes further decrease of the time TV. This is important
for the appearance of the superfluid properties of such a system.
The quasicondensate formation in a nonuniform case is accompanied
by a sharp increase of density. This leads inevitably to the strong
increase of the rate of any inelastic process which depends nonlinearly
on the density [29,7]. In a recent paper [30], it was shown that this fact
can prevent the nonequilibrium system from penetrating deeply into the
Bose condensation region. (It is useful to mention that all the results of
the preceding sections are correct at T ~ Tc.) One can conjecture that
the experimental observation [2] that the excitonic gas does not cross the
224 Yu. Kagan
BEC line while cooling is a consequence of just this effect. The motion
of the system along the BEC line during the evolution, which was also
revealed, can be explained by the adiabatic expansion of the Bose gas
which has the ratio cp/cv = 5/3.
It is clear that the general trends for nonuniform systems can be un-
derstood using the results obtained for the uniform case. However, the
precise values of the relaxation times are very sensitive to the details of
the experimental situation.
References
[I] TJ. Greytak and D. Kleppner, in: New Trends in Atomic Physics, G.
Grynberg and R. Stora, eds. (North-Holland, Amsterdam, 1984) v.2,
p. 1127; I.F. Silvera and J.T.M. Walraven, Progress in Low Temper-
ature Physics, D.F. Brewer, ed. (North-Holland, Amsterdam, 1986),
v.10, p.39; I.F. Silvera and M. Reynolds, Journ. Low Temp. Phys.
87, 343 (1992); see articles by Silvera and by Greytak, this volume.
[2] D.W. Snoke, J.P. Wolfe, and A. Mysyrowicz, Phys. Rev B 39, 11171
(1990); A.Mysyrowicz, D.W. Snoke, and J.P. Wolfe, Phys. Stat. Sol.
159, 387 (1990); see article by Wolfe et al, this volume.
[3] M. Hasuo, N. Nagasawa, T. Itoh, and A. Mysyrowicz, Phys. Rev.
Lett. 70, 1303 (1993).
[4] C. Monroe, W. Swann, H. Robinson, and C. Wieman, Phys. Rev.
Lett. 65, 1571 (1990); see article by Castin et al, this volume.
[5] Yu. Kagan, B.V. Svistunov, and G.V. Shlyapnikov, Pis'ma Zh. Eksp.
Teor. Fiz. 42, 169 (1985) [JETP Lett. 42, 209 (1985)].
[6] Yu. Kagan, B.V. Svistunov, and G.V. Shlyapnikov, Zh. Eksp. Teor.
Fiz. 93, 552 (1987) [Sov. Phys. JETP 66, 314 (1987)].
[7] Yu. Kagan and G.V. Shlyapnikov, Phys. Lett. A130, 483 (1988).
[8] H.E. Hess, G.P. Kochanski, J.M. Doyle, N. Masuhara, D. Kleppner
and TJ. Greytak. Phys. Rev. Lett. 59, 672 (1987); R. van Roijen,
J.J. Berkhout, S. Jakkola, and J.T.M. Walraven, Phys. Rev. Lett. 61,
931 (1988).
[9] E. Fortin and A. Mysyrowicz, Phys. Rev. Lett. 70, 3951 (1993).
[10] E. Levich and V Yakhot, Phys. Rev. B 15, 243 (1977).
[II] P. Nozieres (unpublished).
Kinetics of BEC Formation in an Interacting Bose Gas 225
[12] S.G. Tikhodeev, Zh. Eksp. Teor. Fiz. 97, 681 (1990).
[13] E. Levich and V. Yakhot, J. Phys. A 11, 2237 (1978).
[14] D. Snoke and J.P. Wolfe, Phys. Rev. B 39, 4030 (1989).
[15] H.T.C. Stoof, Phys. Rev. Lett. 66, 3148 (1991).
[16] Yu. Kagan, B.V. Svistunov, and G.V. Shlyapnikov, Zh. Eksp. Teor.
Fiz. 101, 528 (1992)[ Sov. Phys. JETP 75, 387 (1992)].
[17] J.S. Langer, Phys. Rev. 167, 183 (1968).
[18] Yu. Kagan and B.V. Svistunov, Zh. Eksp. Teor. Fiz. 105, No. 2
(1994).
[19] C.P. Barenhi, R J . Donnelly, and W.F. Vinen, Journ. Low Temp.
Phys. 52, 189 (1983).
[20] J.T. Tough, in Progress in Low Temperature Physics, D.F. Brewer,
ed. (North-Holland, Amsterdam, 1982), v.8, p.133.
[21] E.M. Lifshitz and L.P. Pitaevskii, Statistical Physics, Part 2 (Perga-
mon, Oxford, 1980).
[22] V.E. Zakharov, in Basic Plasma Physics Vol. 2, A.A. Galeev and
R.N. Sudan, eds. (North-Holland, Amsterdam, 1984).
[23] B.V. Svistunov, J. Mosccow Phys. Soc. 1, 363 (1991).
[24] L.D. Landau and E.M. Lifshitz, Statistical Physics, Part 1 (Perga-
mon, Oxford, 1980), p. 476.
[25] K.W. Schwarz, Phys. Rev. B 18, 245 (1978).
[26] K.W. Schwarz, Phys. Rev. B 38, 2398 (1988).
[27] S.V. Iordanskii, Zh. Eksp. Teor. Fiz. 49, 225 (1966)[ Sov. Phys. JETP
22, 160, (1966)].
[28] L.D. Landau and E.M. Lifshitz, Fluid Mechanics (Pergamon, Oxford,
1959).
[29] V.V. Goldman, I.F. Silvera, and A.J. Leggett, Phys. Rev. B 24, 2870
(1981).
[30] T.W. Hijmans, Yu. Kagan, G.V. Shlyapnikov, and J.T.M. Walraven,
Phys. Rev. B 48, 12886 (1993); see also this volume.
11
Condensate Formation in a Bose Gas
H. T. C. Stooff
Department of Physics, University of Illinois at Urbana-Champaign
Urbana, Illinois 61801, USA
and
Department of Theoretical Physics, Eindhoven University of Technology
5600 MB Eindhoven, The Netherlands
Abstract
Using magnetically trapped atomic hydrogen as an example, we investigate
the prospects of achieving Bose-Einstein condensation in a dilute Bose gas.
We show that, if the gas is quenched sufficiently far into the critical region
of the phase transition, the typical time scale for the nucleation of the
condensate density is short and of O(h/kBTc). As a result we find that
thermalizing elastic collisions act as a bottleneck for the formation of the
condensate. In the case of doubly polarized atomic hydrogen these occur
much more frequently than the inelastic collisions leading to decay and we
are led to the conclusion that Bose-Einstein condensation can indeed be
achieved within the lifetime of the gas.
1 Introduction
In the last few years it has been clearly demonstrated that not only
charged ions but also neutral atoms can be conveniently trapped and
cooled by means of electro-magnetic fields. Although the physics of
the various ingenious scenarios developed to accomplish this is already
interesting in itself [1], the opportunities offered by an atomic gas sample
at very low temperatures are exciting in their own right. Examples in this
respect are the performance of high-precision spectroscopy, the search
for a violation of CP invariance by measuring the electric dipole moment
of atomic cesium [2], the construction of an improved time standard
based on an atomic fountain [3] and the achievement of Bose-Einstein
condensation in a weakly interacting gas.
In particular the last objective has been pursued mainly with atomic
t Present address: Institute for Theoretical Physics, University of Utrecht, Utrecht,
The Netherlands.
226
Condensate Formation in a Bose Gas 227
hydrogen [4, 5]. It has been recently proposed that the alkali-metal
vapors cesium [6] and lithium [7] are also suitable candidates for Bose-
Einstein condensation. We will nevertheless concentrate here on atomic
hydrogen, because it still seems to be the most promising system for
the observation of the phase transition in the near future. (See the
review articles by Greytak and Silvera in this volume.) Moreover, it has
the advantage that the atomic interaction potential is known to a high
degree of accuracy. As a result we can have confidence in the fact that
the scattering length is positive, which is required for the condensation to
take place in the gaseous phase [8], and small enough to rigorously justify
the approximations made in the following for the typical temperatures
(T ~ 10 JUK) and densities (n ~ 1014 cm"3) envisaged in the experiments.
Due to the spin of the electron and the proton, the ls-hyperfine
manifold of atomic hydrogen consists of four states which are in order of
increasing energy denoted by \a), \b), \c)9 and \d), respectively. Only the
\c) and \d) states can be trapped in a static magnetic trap, because in a
magnetic field they have predominantly an electron spin-up component
and are therefore low-field seeking [9]. Furthermore, if we load a trap with
atoms in these two hyperfine states, the \c) state is rapidly depopulated
as a result of the much larger probability for collisional relaxation to the
high-field seeking \a) and \b) states which are expelled from the trap. In
this manner the system polarizes spontaneously and we obtain a gas of
|d) -state atoms, known as doubly spin-polarized atomic hydrogen since
both the electron and the proton spin are directed along the magnetic
field. Unfortunately, such a doubly polarized hydrogen gas still decays
due to the dipole interaction between the magnetic moments of the
atoms. Although the time scale Tinei for this decay is much longer than
the time scale for the depopulation of the \c) state mentioned above, it
nevertheless limits the lifetime of the gas sample to the order of seconds
for the densities of interest [10].
Having filled the trap with doubly polarized atoms, we must subse-
quently lower the temperature of the gas to accomplish Bose-Einstein
condensation. At present it is believed that the most convenient way to
achieve this is by means of conventional [11, 12] or light-induced [13]
evaporative cooling. In both cases the idea is to remove, by lowering
the well-depth or by photon absorption in the perimeter, the most en-
ergetic particles from the trap and thus to create momentarily a highly
nonequilibrium energy distribution that will evolve into a new equilib-
rium distribution at a lower temperature. According to the quantum
Boltzmann equation describing this process, a typical time scale for the
228 H. T. C. Stoof
evolution is the average time between two elastic collisions rei = l/n(v(r),
with (vo) the thermal average of the relative velocity v of two colliding
atoms times their elastic cross section a. Clearly, rei must be small com-
pared to Tinei to ensure that thermal equilibrium is achieved within the
lifetime of the system. As a result, the minimum temperature that can
be reached by evaporative cooling is about 1 /xK and indeed below the
critical temperature of atomic hydrogen at a density of 1014 cm"3.
The previous discussion appears to indicate that a typical time scale
for the formation of the condensate is given by rei. However, this is
not correct because simple phase-space arguments show that a kinetic
equation cannot lead to a macroscopic occupation of the one-particle
ground state: considering a homogeneous system of N bosons in a
volume V, we find from the Boltzmann equation that the production
rate of the condensate fraction is
d No
(1)
dt N
where No is the number of particles in the zero-momentum state and C
is a constant of 0(1). Hence, in the thermodynamic limit (N, V —• oo in
such a way that their ratio n = N/V remains fixed) a nonzero production
rate is only possible if a condensate already exists [14] and we are forced
to conclude that Bose-Einstein condensation cannot be achieved by
evaporative cooling of the gas.
2 Nucleation
In the above argument we have only considered the effect of two-body
collisions. It is therefore legitimate to suggest that perhaps three- or
more body collisions are required for the formation of the condensate,
even though they are very improbable in a dilute gas [15]. However, we
can easily show that the same argument also applies to these processes:
in a m-body collision that produces one particle with zero momentum,
we have 2m —2 independent momentum summations, leading to a factor
of V2m~2. Moreover, the transition matrix element is proportional to
V - V~m due to the integration over the center-of-mass coordinate and the
normalization of the initial and final state wave functions, respectively. In
total, the production rate for the condensate fraction is thus proportional
to vlm-\Vx-m)2V-\\ +N0) or V-^l+No), which again vanishes in the
thermodynamic limit if there is no nucleus of the condensed phase. As
expected, the contributions from collisions that produce more than one
Condensate Formation in a Bose Gas 229
zero-momentum particle have additional factors of V~l and vanish even
more rapidly if V —> oo.
Clearly, we have arrived at a nucleation problem for the achievement
of Bose-Einstein condensation which seriously endangers the success of
future experiments. Fortunately, we suspect that the line of reasoning
presented above is not completely rigorous because otherwise it implies
that liquid helium also cannot become superfluid, in evident disagreement
with experiment. Indeed, by using a kinetic equation to discuss the time
evolution of the gas we have in effect neglected the buildup of coherence
which is crucial for the formation of the condensate. Our previous
argument therefore only shows that by means of evaporative cooling
the gas is quenched into the critical region on a time scale xei, not
that Bose-Einstein condensation is impossible. To discuss that point we
need a different approach that accurately describes the time evolution of
the system after the quenching by taking the buildup of coherence into
account exactly. Such a nonequilibrium approach was recently developed
on the basis of the Keldysh formalism and can, in the case of a dilute
Bose gas, be seen as a generalization of the Landau theory of second-
order phase transitions [16, 17]. As a consequence it is useful to consider
the Landau theory first. This leads to a better understanding of the
more complicated nonequilibrium theory and ultimately of the physics
involved in the nucleation of Bose-Einstein condensation.
where J is the exchange energy and the sum is only over nearest neighbors.
For further convenience we also introduce the magnetization
for small values of (M), and that the coefficients of this expansion behave
near the critical temperature as
( ^ ) (5)
and
pm^Po, (6)
respectively, with ao and /?o positive constants.
Hence, above the critical temperature a(T) and fi(T) are both positive.
As a result the free energy is minimal for (M) = 0, which corresponds
exactly to the paramagnetic phase. Moreover, for temperatures below the
critical one a(T) is negative and the free energy is indeed minimized by
a nonzero average magnetization with magnitude y/—ot(T)/f}(T). Just
below the critical temperature the latter equals
(V)
> T ) / ( 0 , T ) ^ ( l (8)
[a)
Fig. 1. Visualization of (a) the time scale rCOh for the relaxation of the order
parameter to its equilibrium value and (b) the time scale Tnuci associated with the
appearance of the instability.
before this can be done we need to know the correct order parameter of
the phase transition.
(b)
Fig. 2. Visualization of the time scales iCOh and Tnuci, using the time dependence
of (a) the order parameter and (b) the coefficient a of the quadratic term in the
free energy.
N = [ dx\p\x)\p{x). (10)
The method is also particularly useful for a Bose system because the per-
mutation symmetry of the many-body wave function is automatically ac-
counted for by assuming the Bose commutation relations [\p(x),\p(xf)] =
[xp^x^xp^x')] = 0 and [t/^xXt/^x')] = d(x — xf) between the creation
and annihilation operators.
234 H. T. C. Stoof
In the language of second quantization, the order parameter for the
dilute Bose gas is the expectation value (ip(x)). Analogous to the case
of the ferromagnetic phase transition, a nonzero value of this order pa-
rameter signals a spontaneously broken symmetry. Here the appropriate
symmetry group is (7(1), since the Hamiltonian of Eq. (9) is invariant
under the transformation xp(x) —• xp(x)eiS and y^(x) —• xp^(x)e~i9 of the
field operators, whereas their expectation values are clearly not. Notice
that the (7(1) symmetry of the Bose gas is closely related to the conser-
vation of particle number. This is most easily seen by observing that
the invariance of the Hamiltonian is due to the fact that each term in
the right-hand side of Eq. (9) contains an equal number of creation
and annihilation operators. The relationship can also be established in
a more formal way by noting that the (7(1) gauge transformations are
generated by the particle number operator. As we will see later on, it has
important consequences for the dynamics of the order parameter.
To understand why (ip(x)) is the order parameter associated with
Bose-Einstein condensation, it is convenient to use a momentum-space
description and to introduce the annihilation operator for a particle with
momentum hk
(11)
and we can neglect the fact that ao and a^ do not commute. As a result
we can treat these operators as complex numbers [18, 19] and say that
(OO^OQ) = (ao*)(ao) or equivalently that (ao) = y/No- In coordinate space,
the latter reads (ip(x)) = Jno, with no = No/V the condensate density.
The above argument essentially tells us that a sufficient condition for
a nonzero value of (xp(x)) is (No) > 1. Although this is intuitively
appealing, it is important to point out that it is not generally true.
Consider, for example, the ideal Bose gas [20]. In the grand canonical
Condensate Formation in a Bose Gas 235
ensemble, the total number of particles in the gas is given by
^ (14)
Lowering the temperature while keeping the density fixed, the fugacity
increases until it ultimately reaches the value one at the critical temper-
ature
T =
2 (n ^ ( n \ 2 n 2*tf/nx2/3
° mkB Vg3/2(1)/ " mkB
At this point, Eq. (14) ceases to be valid because the occupation number
of the zero-momentum state, which is equal to £/(l — f), diverges and
must be taken out of the discrete sum in Eq. (13) before we take the
continuum limit. Moreover, we only need to treat the zero-momentum
term separately because in the thermodynamic limit the chemical po-
tential goes to zero as F" 1 , whereas the kinetic energy for the smallest
nonzero momentum decreases only as F~ 2/3 . Consequently, below the
critical temperature the equation of state becomes
ll = Mx,t),H] (20)
Condensate Formation in a Bose Gas 237
for the field operator. Substituting the Hamiltonian in Eq. (9), and taking
the expectation value with respect to p(to\ we find
2
) V(*,t)>, (24)
having only the trivial solution (xp(\9t)) = 0 for a space and time-
independent order parameter. Within this lowest order approximation
we thus conclude that Tnuci = oo and that the formation of a condensate
will not take place.
Fortunately, it is well known that the Hartree-Fock approximation
is not sufficiently accurate for a dilute Bose gas because the diluteness
condition na3 <C 1 implies that we need to consider all two-body pro-
cesses, i.e., two particles must be allowed to interact more than once. The
appropriate approximation is therefore the ladder or T-matrix approx-
imation, which is displayed in terms of diagrams in Fig. 3. Moreover,
in the degenerate regime, where the temperature T is slightly larger
than To and the degeneracy parameter nA3 is of 0(1), the condition
238 H. T. C. Stoof
a/A <C 1 implies that also naA2 < 1 or physically that the average ki-
netic energy of the gas is much larger than the typical interaction energy.
Consequently, an accurate discussion of the nucleation of Bose-Einstein
condensation in a weakly interacting Bose gas requires an evaluation of
(v?t(x', t)xp(x\ t)xp(x, t)) within the T-matrix approximation and to zeroth
order in the gas parameters a/A and naA2.
Although it is easy to formulate this objective, to actually perform the
calculation is considerably more difficult. It is most conveniently accom-
plished by making use of the Keldysh formalism [22] which has been
reviewed by Danielewicz [23] using operator methods. For a functional
formulation of this nonequilibrium theory and for the technical details of
the somewhat tedious mathematics, we refer to our previous papers [16].
Here we only present the final results and concentrate on the physics
involved.
Due to the fact that we are allowed to neglect the (relative) momentum
dependence of the T matrix, the equation of motion for the order
parameter (tp(x, t)) acquires the local form of a time-dependent Landau-
Ginzburg theory
^d h2V2
ih (xp(x,t)) = , (25)
Jt
s(+
Fig. 4. Time dependence of the coefficient for three different initial temper-
atures of the Bose gas.
S((xp(x,t)),T) = J dt J dx{ip(x,t)y
x in— . (26)
(£ j* j (28)
Substituting the latter into the action of Eq. (26) and minimizing with
respect to jno(t) gives, for t > tC9
n
o(t) =
plane, (y>(x, t)) thus moves radially outward along a spiral, as shown in
Fig. 5. Consequently, the phase of the order parameter never has a fixed
value and the U(l) symmetry is not really broken dynamically. This is
of course expected since the system evolves according to a symmetric
Hamiltonian.
(31)
) v(x,t)), (32)
which has the equilibrium solution (tp(x, t)) = Jn$ exp(—i^oi) and /zo =
noT (+) . Consequently, all the particles that initially have momenta hk <
hko are in the limit t —• oo assumed to be in the condensate.
Condensate Formation in a Bose Gas 243
Linearizing the Hamiltonian (32) around this equilibrium solution,
Kagan et a\. then observe that the energy involved with a magnitude
fluctuation of the order parameter is e^ + noT(+), whereas the energy
involved with a phase fluctuation is only e^. As a result, they assert
that on the time scale Tampi = Tcoh = 0(h/noT^), a state is formed in
which the amplitude of (xp(x,t)) is fixed, but the phase is still strongly
fluctuating because the corresponding time scale xph is much longer (and
even diverges as V2^ in the thermodynamic limit). Hence, at finite time,
the gas is in a state with only a quasicondensate [25, 26] and a complete
or true condensate is only formed in the limit t —• oo.
Clearly, this physical picture of two different time scales for the am-
plitude and phase fluctuations of the order parameter is only applicable
if these fluctuations exist independently of each other. Looking only at
the Hamiltonian this indeed seems to be the case. However, a correct
discussion of the fluctuations must be based on the equations of mo-
tion or, equivalently, the Lagrangian. The latter contains a first-order
time derivative which strongly couples the amplitude and phase fluctu-
ations. Therefore, a dilute Bose gas only has one collective mode (not
two) with a single dispersion relation, i.e. the well-known Bogoliubov
dispersion \A"k(ek + 2YIQT^\ and we are lead to TPH = Tampi- It is in-
teresting to note that in the case of a neutral BCS-type superfluid, we
do have two different time scales because the Lagrangian now contains
a second-order time derivative and the amplitude and phase fluctuations
are indeed independent in lowest order [27, 28].
An even more serious problem with the approach of Kagan, Svistunov
and Shlyapnikov is their claim that the use of the initial condition in Eq.
(31) is justified because (N^) > 1. As we have pointed out earlier, this is
not true in general. For (\p(x91)) to be nonzero, one must show that the
system has a corresponding instability. However, within the T -matrix
approximation it is not difficult to show that the instability associated
with a quasicondensate is always preceded by the instability correspond-
ing to the formation of a condensate. This implies that we always have to
take Bose-Einstein condensation into account first. After that has been
accomplished by means of the theory reviewed here, it is of course no
longer relevant to consider the appearance of a quasicondensate.
References
[I] See the special issue of J. Opt. Soc. Am. B 6, No. 11 (1989), edited
by S. Chu and C. Wieman.
[2] W. Bernreuther and M. Suzuki, Rev. Mod. Phys. 63, 131 (1991).
[3] A. Clairon, C. Salomon, S. Guellati, and W.D. Phillips, Europhys.
Lett. 12, 683 (1990); K. Gibble and S. Chu, Phys. Rev. Lett. 70,
1771 (1993).
[4] H.F. Hess, G.P. Kochanski, J.M. Doyle, N. Masuhara, D. Kleppner,
and TJ. Greytak, Phys. Rev. Lett. 59, 672 (1987); N. Masuhara,
J.M. Doyle, J.C. Sandberg, D. Kleppner, T.J. Greytak, H.F. Hess,
and G.P. Kochanski, Phys. Rev. Lett. 61, 935 (1988).
[5] R. van Roijen, J.J. Berkhout, S. Jaakkola, and J.T.M. Walraven,
Phys. Rev. Lett. 61, 931 (1988).
[6] C. Monroe, W. Swann, H. Robinson, and C. Wieman, Phys. Rev.
Lett. 65, 1571 (1990).
[7] J.J. Tollett, C.C. Bradley, and R.G. Hulet, Bull. Am. Phys. Soc. 37,
1126(1992).
[8] H.T.C. Stoof, to be published.
[9] For a review we refer to T.J. Greytak and D. Kleppner, in New
Trends in Atomic Physics, edited by C. Grynberg and R. Stora
(North-Holland, Amsterdam, 1984), p. 1125 and I.F. Silvera and
J.T.M. Walraven, in Progress in Low Temperature Physics, edited by
D.F. Brewer (North-Holland, Amsterdam, 1986), Vol. 10, p. 139.
[10] A. Lagendijk, I.F. Silvera, and B.J. Verhaar, Phys. Rev. A 33, 626
(1986); H.T.C. Stoof, J.M.V.A. Koelman, and B.J. Verhaar, Phys.
Rev. B 38, 4688 (1988).
[II] J.M. Doyle, J.C. Sandberg, LA. Yu, C.L. Cesar, D. Kleppner, and
TJ. Greytak, Phys. Rev. Lett. 67, 603 (1991).
[12] O J . Luiten, H.G.C. Werij, I.D. Setija, T.W. Hijmans, and J.T.M.
Walraven, Phys. Rev. Lett. 70, 544 (1993).
[13] I.D. Setija, H.G.C. Werij, O J . Luiten, M.W. Reynolds, T.W. Hijmans,
and J.T.M. Walraven, Phys. Rev. Lett. 70, 2257 (1993).
[14] See also E. Levich and V. Yakhot, Phys. Rev. B 15, 243 (1977) and
S.G. Tikhodeev, Zh. Eksp. Teor. Fiz. 97, 681 (1990) [Sov. Phys.-JETP
70, 380 (1990)].
[15] D.W. Snoke and J.P. Wolfe, Phys. Rev. B 39, 4030 (1989).
Condensate Formation in a Bose Gas 245
[16] H.T.C. Stoof, Phys. Rev. Lett. 66, 3148 (1991); Phys. Rev. A 45,
8398 (1992).
[17] For a more detailed discussion see, for instance, S. Ma, Modern
Theory of Critical Phenomena (Addison-Wesley, New York, 1976)
and J.W. Negele and H. Orland, Quantum Many-Particle Systems
(Addison-Wesley, New York, 1988).
[18] N.N. Bogoliubov, J. Phys. USSR (Moscow) 11, 23 (1947).
[19] See, for example, A.L. Fetter and J.D. Walecka, Quantum Theory of
Many-Particle Systems (McGraw-Hill, New York, 1971).
[20] K. Huang, Statistical Mechanics, second edn (Wiley, New York,
1987).
[21] N.M. Hugenholtz and D. Pines, Phys. Rev. 116, 489 (1958).
[22] L.V. Keldysh, Zh. Eksp. Teor. Fiz. 47, 1515 (1964) [Sov. Phys.-JETP
20, 235 (1965)].
[23] P. Danielewicz, Ann. Phys. (N.Y) 152, 239 (1984).
[24] U. Eckern, J. Low Temp. Phys. 54, 333 (1984).
[25] Yu. M. Kagan, B.V. Svistunov and G.V. Shlyapnikov, Zh. Eksp.
Teor. Fiz. 101, 528 (1992) [Sov. Phys.-JETP 74, 279 (1992)]; Sov.
Phys.-JETP 75, 387 (E) (1992); B.V. Svistunov, J. Moscow Phys. Soc.
1, 373 (1991); Yu. Kagan, this volume.
[26] V.N. Popov, Functional Integrals in Quantum Field Theory and Sta-
tistical Physics (Reidel, Dordrecht, 1983).
[27] N.N. Bogoliubov, V.V. Tolmachev, and D.N. Shirkov, New Methods
in the Theory of Superconductivity (Consultants Bureau Enterprises,
New York, 1959).
[28] RW. Anderson, Phys. Rev. 112, 1900 (1958).
12
Macroscopic Coherent States of Excitons
in Semiconductors
L. V. Keldysh
P. N. Lebedev Physics Institute
Leninsky Prospect 53
117924 GSP, Moscow B-333
Russia
Abstract
Initially put forward by Moskalenko [1] and Blatt et al. [2], the idea of a
possible Bose-Einstein condensation (BEC) of excitons in semiconductors
has attracted the attention of both experimentalists and theoreticians for
more than three decades. At different stages of this long history, the results
of their efforts have been described and discussed in review articles [3-10].
A brief introduction and summary of the main qualitative conclusions of
this older work is presented here (Sections 1 and 2), followed by a more
detailed discussion of some more recent developments (Sections 3 and 4).
246
Macroscopic Coherent States of Excitons in Semiconductors 247
-
Conduction band
• e
'-' cmin
Oh
Oh
Valence band
is a few orders of magnitude smaller than the atomic binding energy, and
the exciton effective radius
eh2
a0 = —- 2~ 10" 7 - l(T6cm (2)
me
is macroscopically large compared to interatomic distances in the host
crystal. Formulas (1) and (2) represent the natural units of quantum
scales of energy and length in the interacting electron-hole system, in
just the same sense as the Rydberg and Bohr radius are the natural scales
in atomic, molecular and solid state physics. Therefore in what follows
we will use these units. Instead of the temperature, a dimensionless
ratio kT/Eo, designated by T, will be used, and instead of particle
concentration the product na\, designated by n. In this notation, the
usual formula for the critical temperature of an excitonic gas BEC,
neglecting any interaction corrections, is given by
TC = 6.62^(V 3 . 0)
M g
Here M = me + m^ is the effective mass of the exciton, equal to the
sum of electron me and hole nih effective masses; m — m e- nih/M is their
reduced mass, entering (1) and (2); and g is the exciton ground state
statistical weight (degeneracy).
(2) The effective masses me and nth in any particular semiconductor,
though different, are usually of the same order of magnitude (or differ
at most by one order of magnitude). The absence in the electron-hole
system of really heavy particles such as nuclei results in a dominant role
of quantum effects at all temperatures T < 1. In particular as formula
(3) shows, Tc ^ 1 for n ~ 1, unlike, for example, in 4He where Tc ~ 10~5
in corresponding units. Other important manifestations of the absence of
heavy particles are very large zero vibrations of excitons in the excitonic
molecule. The amplitude of this vibration is of the order of, or even
larger than, the excitonic radius CIQ itself. So biexcitons appear to be
Macroscopic Coherent States of Excitons in Semiconductors 249
very loosely bound complexes with binding energy ED not exeeding 0.1,
compared to 0.35 in the hydrogen molecule [16, 17].
For the same reason, nothing like an "excitonic crystal" can exist. The
weak van der Waals attraction dominating at large intermolecular dis-
tances is not able to confine light particles such as excitons or biexcitons.
According to Ref. [18], the s-wave scattering length of two excitonic
molecules is large (~ 7) and positive (repulsive). Therefore a condensed
phase of excitonic molecules - a molecular "liquid" similar to liquid
hydrogen - also cannot exist. But an electron-hole liquid (EHL) similar
to metallic hydrogen or alkali metals does exist [15, 6, 9, 10]. Unlike
common metals, in the EHL not only electrons but also holes ("nuclei")
are Fermi degenerate. If the effective masses of both electrons and holes
are more or less isotropic, the EHL at low temperatures transforms to
an "excitonic insulator" phase [19-21]. In this two-Fermi-liquid state the
collective pairing of electrons and holes in the vicinity of Fermi surfaces
arises, quite similar to BCS pairing in superconductors. This pairing
manifests itself in the appearance of the energy gaps A around the Fermi
surfaces, as shown in Fig. 2. These gaps may be considered a rem-
nant of the binding energy of a single exciton, transformed by collective
many-body interactions in the condensed phase. The gap diminishes fast
with increasing particle density or me-nth difference, and it does not exist
at all if the anisotropy is large. In that sense, the excitonic insulator
state in the nonequilibrium electron-hole (e-h) system is a coherent BEC
state of high density (n > 1) excitons, as the superconducting state is a
coherent BEC state of Cooper pairs. Like Cooper pairs, e-h pairs in an
excitonic insulator have very large radii ~ A"1/2, much larger than in-
terparticle distance. So they are true collective phenomenon. But, unlike
Cooper pairs in superconductors which are charged (—2e), excitons are
electrically neutral.
The above-described phenomena can be schematically represented by
the phase diagrams depicted in Figs. 3-4. They show possible states of
the e-h system in the plane (n, T). Here ft is the average particle density.
The phase diagram type of any particular semiconductor is defined
by the relationship of the binding energy per e-h pair in the biexciton
gas, given by (Eo + ED/2)9 and in the EHL, indicated by —/x/. Strictly
speaking, \i\ is the difference between the chemical potential per e-h pair
in the EHL and £ g , where Eg is the minimum energy of the free e-h
pair. Different phases presented in Figs. 3 and 4 are: G, the gas (plasma)
of weakly interacting electrons, holes, excitons and excitonic molecules;
BG, the degenerate gas of excitons and(or) excitonic molecules; ML, the
250 L. V. Keldysh
metallic liquid (EHL); and IL, the insulating liquid (excitonic insulator).
The striped area is the region of phase coexistance: droplets of liquid
phase surrounded by gas, degenerate or nondegenerate. Coherent states
BG or IL are present in the phase diagrams of the types shown in Figs.
3(b) and 4(a),(b). But up to now only the simplest type, shown in Fig.
3(a), is firmly established and studied experimentally, in germanium and
silicon.
Both phases in Fig. 3 correspond to the case | /x/ | > (EQ + ED/2);
that is, the EHL is more tightly bound than excitonic molecules. It
is theoretically proven and experimentally confirmed that the fulfilment
of this condition is strongly favored by multivalley band structure, i.e.
the presence in the electronic spectrum of a semiconductor of a few
equivalent groups of electrons or holes (or both). Just such a band
structure is characteristic of both germanium and silicon. No BEC exists
in the gas phase in this case, exactly as in the case of helium. It does exist
in the phase diagrams depicted in Fig. 4, corresponding to the fulfillment
Macroscopic Coherent States of Excitons in Semiconductors 251
ML
ne(T)
log
(a)
Fig. 3. Possible types of nonequilibrium electron-hole system phase diagrams in
semiconductors with electron-hole liquid more tightly bound than the excitonic
molecule. The striped regions are the coexistence domains of the liquid and gas
phases.
o ne|
(a) (b)
Fig. 4. The same as in Fig. 3, but for semiconductors with excitonic molecules
more tightly bound than electron-hole liquid.
where cpo is the wave function describing the relative e-h motion in the
exciton. The commutator of the two exciton operators, AF and A?>, is
then
"P' "P
P . "I
(8)
where Q = (P' - P)/2. The righthand side of (8) differs from the usual
one for Bose operators <5pfp by operators that are of the order of n, the
exciton density. So for n ~ 1, the operator (7) has nothing to do with usual
254 L. V. Keldysh
2 Polaritons
A very peculiar, but undoubtedly also the most important, case is pre-
sented by BEC of dipole-active excitons. Not only is it always essentially
a nonequilibrium problem, as explained above, but the nature of the
quasiparticles themselves is very special. They are the well-known polari-
tons - the quantum superposition of photons and electronic excitation
(excitons) [25]. In other words, a polariton is a real photon in a medium:
an exciton representing a polarization cloud accompanying propagation
of electromagnetic field through the medium [26]. These polarization ef-
fects are especially strong under resonant conditions, that is, for photon
energies hco close to the energy of exciton creation ha>o = (Eg — Eo).
Schematically depicted in the Fig. 6 is the well-known polariton disper-
sion law. Strong mixing arises close to the intersection point of the
undisturbed photon and exciton branches (ko,&>o), and the typical pic-
ture of level repulsion appears. So the problem of BEC of dipole-active
excitons is essentially that of photons, or, in other words, the problem of
a coherent electromagnetic wave of finite amplitude in the polariton fre-
quency range. It is a very familiar problem but is still far from trivial. It
includes all the many-body phenomena typical of BEC, but they acquire
the meaning of nonlinear optics phenomena. Two different problems are
usually considered in this context.
The accumulation of a macroscopic number of initially incoherent ex-
citation quanta in a single-photon mode is lasing. It was recognized long
ago [27-31] that the appearance of a coherent mode from multimode
intense noise in a pumped system is a very typical example of a nonequi-
librium phase transition. The approaches used for its general theoretical
treatment are based on mean-field approximations, the Langevin equa-
tion, the Fokker-Planck equation, etc. Their detailed review is beyond
the scope of this article. As applied to the particular case of dipole-active
excitons in semiconductors, the qualitative picture of the lasing process
should look like the following [32]. Initially excited electrons and holes
combine into excitons, which dissipate their kinetic energy by emission
of phonons - quanta of the crystal lattice vibrations - and so move
downwards in energy along the lower polariton branch (LPB) in Fig. 6.
Approaching the exciton-photon intersection point, which corresponds
to very small momenta ~ Eg/c (c the light velocity), they smoothly
256 L. V. Keldysh
1 1
/ /
UPB
.. — //^ LPB
////
//
//
////
i i
where
p=- (10)
EE
is the matrix element of the electronic dipole transition between the
conduction and valence bands induced by the electric field S\ po is the
momentum of the photons; and p w is the interband momentum matrix
element. In its physical interpretation, / is the interband Rabi frequency
in terms of Eo/h.
It was indicated in Ref. [34] that for photon energies above the ab-
sorption threshold hco > Eg, a coherent field produces a collective excited
state of semiconductors quite similar to the excitonic insulator state, even
neglecting the Coulomb interaction. Creating electrons and holes coher-
ently in correlated pairs, the field itself acts as the pairing force. Also,
an energy gap arises in the electronic spectrum at the Fermi level with a
magnitude | / | (Fig. 2). The obvious condition for the existence of such
a state is / T > 1, where T is electron (hole) relaxation time. The complete
treatment of this problem needs inclusion of many-body effects. This was
done in Refs. [35-36] (see also the reviews and monographs [31, 33], [38-
40] and especially [37], where the problem has received its most complete
formulation, in terms of nonequilibrium Green functions [41]. (The de-
258 L. V. Keldysh
tails of this technique are presented in Refs. [42-51], [31] and many
other reviews and monographs.) Here we summarize only a few general
definitions and basic relations and also some features specific to the
nonequilibrium e-h system.
The nonequilibrium Green's function is a 2 x 2 matrix, which in the
so-called "triangular representation" looks as follows:
0 G \
) <»>
Here x = (r, t)\ indices +(—) in (12) and —(+) in (14) refer to the case of
Fermi (Bose) statistics. In the electron-hole system all of these functions
are 2 x 2 matrices in band indices (c9v). In this notation, the interband
(electron-hole) Green's function
= -f<[^(x),^V)]_>. (17)
= 5ijd(x-x'l (19)
where "self-energy" matrices %j also have "triangular" form
(2o)
*•• - (% f )
and can be expressed in terms of Green's functions and the interaction
potential in any approximation by a set of Feynman graphs. The specific
feature of the system under consideration is that, apart from the infinite
set of diagrams accounting for different order interaction processes, the
interband self-energy Ee/, contains, according to (9), the driving force
contribution
f<o)/Y Y/x = ( 0 f{x
ij( }
* \f(x) 0
The electric field $ entering into Eq. (10) has to be found self-consistently
from Maxwell's equation,
(22)
Sjy ~ d(t — tf), which is the case at least in the mean-field approximation,
260 L. V. Keldysh
(19) can be reduced to the equation for the single-time (t = tr) density
matrix [31, 33, 36, 37, 51]
= (k + i ) f f ( h o + ± ) . (25)
Here the two-band Hamiltonian matrix is introduced by
fuu - f(°) ^ ^ 1 ^h i
Uhh
~ feM £(P) + z«(p
£ S(r)
( (P)
P) + (P) ,f(Q)
(Q) , ,(0k
(0k ,28v
(28)
where
4W ^ f'5(r^) (31)
Macroscopic Coherent States of Excitons in Semiconductors 261
and so the set of equations (25)-(30) is complete. Under stationary
conditions they reduce to two equations for the self-energies,
u (32)
_
(P)
-
The most important equation here is (32). It defines the gap in the
renormalized excitation spectrum,
+ T2 (—
\m e ~
For hco0 > \Eg+ Tir\0) | it is just equal to | I,eh |. So (32) is similar
to a BCS equation, but inhomogeneous and with a slightly different
integral kernel. The last difference is due to the nonequilibrium nature
of our problem. At moderate field | / | < 1, and not too close to
the resonance | Z^ | <C Eg — ha>o, Eq. (32) can be transformed to an
inhomogeneous Schrodinger equation for the exciton wave function. At
very large intensities | / | > 1, it gives the gap | Z^ |« /, as was obtained
in Ref. [34]. The functions / $ entering (32)-(33), defined by Eq. (30),
are connected to spontaneous fluctuations of occupation numbers. In
thermodynamic equilibrium, the ratios on the right-hand side of (31) are
found to be exactly
where
0 1
T = l
1 0
264 L. V. Keldysh
G g
a,x 0,x' a,x fiX
w
(p, c) (P, 0 (P, « ) (Po> o)
(38)
J MP;*>0 -
J(p; t, fi)|y/Kp; tu tf)dh = S^S(t - 0 , (40)
266 L. V. Keldysh
at
\ J J
°g ) • (42)
Here, Q ^ is the noise correlator in the system, closely related to the
probability of incoherent creation and annihilation of excitons. The
macroscopic wave function tpa(Po>O of coherent excitons satisfies the
following field equation:
l d ,0A / x f irh ,. , /w/
T+ ~ 6PO V«(Po, t) - / (TjJ(po; t, t ) ^ ( p o ; t )dt
at j j
777/
Fig. 8. Diagonal (E) and anormalous (£) self-energy graphs in the mean-field
approximation.
The equations (40)-(41) are the matrix equations in the sense of the
definitions (38) and (42). In fact, the equations for g (r) and g (r) look
exactly like (40) and (41) with the substitution of the proper self-energy
functions. The same statement holds for the advanced Green's functions.
As to the function f^(p;t,tf), it can be expressed in terms of the self-
energies Q ^ , and the retarded and advanced functions [41]:
(46)
(47)
268 L. V. Keldysh
(48)
4= |2 • (51)
The values ejj1 represent the energy difference between the excitonic levels
and the carrier frequency of the pump. In a sense, the last formula (52)
is an odd result. Instead of the usual resonance relation of the exciton
polarization \p to thefield/, the denominator in Eq. (52) is quite different
from 6jj=. This result is one of the manifestations of the presence of pair
coherence in the system under consideration.
The polarizability %(k, co) which describes the propagation of a weak
probe signal (k,co) in a semiconductor in the presence of a pump is given
by the following expression:
4TT I m + |2| g(k) |2
g(k) 2 (53)
co
CO — c 2 fe 2 — -
This formula directly manifests the interconnection of the polariton
branches k and k. The proper Green's function D(k) of the probe
electromagnetic field is given by
(co - e£)A+ + vp | m+ |2 6+
D(k) = (54)
A~A+ - 2vk | m +
with
A =(co — — e£)— \ m + (55)
A+ = (co - m+ \ (56)
and
Vn = - E — T 1 (57)
Macroscopic Coherent States of Excitons in Semiconductors 269
The poles of D(k) describe the four new polariton-like branches. Each of
these branches reflects the joint propagation of the four particles - two
excitons and two photons with wave vectors k and k. Their dispersion
laws are
Here cffi = c|p| and np = — 0.5 [1 + i/*x(p; U t)] are the exciton occupation
numbers. It is easy to see from (54), and especially (58), that the coupling
of two polaritons p and p is determined by vp, which in its turn is
governed essentially by the | #+ |2? that is, by the EM density.
4 Results
In the framework of the model introduced, the considerations in the
previous section are quite general. In order to obtain a more detailed
description accounting for explicit dependences of self-energies on pump-
ing frequency and intensity, the mean-field approximation will be used
below. In this case a matrix equation similar to (25) can be used for the
density matrix /^(r, r'; t, i):
-dard(r-rf)] (60)
and
h = I - (Eg - Eo - hcoo - ^vA . (61)
2
/
o
o
\ "2-
Fig. 10. The same as in Fig. 9, but for / = 0.25 and a = 0.6.
exists around coo = 0. Only a loop in the interval 0.5 < a>o < 2 in
Fig. 9, also disappearing at higher field, is a small remnant of a "pure"
excitonic resonance observed at low fields. Instead of that, a typical
resonant structure for excitonic molecules is observed, accompanied by a
strongly nonlinear resonance of the excitonic polarization in the vicinity
of COQ = — 1. For frequencies above this stationary resonance the response
Macroscopic Coherent States of Excitons in Semiconductors 273
Fig. 11. The same as in Fig. 9, but for / = 0.25 and a = 0.3.
becomes multistable. Two new states arise. The one with the smaller
value of excitonic amplitude corresponds to the one with larger biexciton
amplitude, and vice versa. Strictly speaking, solutions with coo > 0, c$ < 0
are unstable because of stimulated scattering into some states in the lower
polariton branch. Therefore, they will not be discussed in what follows.
Among the three stable (c0 > 0) solutions, one is "exciton dominated", i.e.
beyond the immediate vicinity of the resonance, the majority of excitons
are in the single-particle condensate-coherent polarization cloud. Two
other solutions are "biexciton dominated" - the majority of excitons
are bound into molecules at the expense of coherent polarization. But
nonlinear processes become greatly enhanced, such as four-wave mixing
(FWM) which increases with a, as can readily be seen from (54)-(57)
and the following formula, relating the exciton distribution function to
the anomalous self-energy
1
(71)
?U-J
Fig. 12. Field dependence of the coherent exciton amplitude Co (full lines; note
the difference with Figs. 9-11!) and anomalous self-energy 2 (broken lines) for
pumping frequency COQ = —0.75.
coshz (72)
1-a
with a finite value of biexciton density | <r( | « sinhz. Here z is the root
Macroscopic Coherent States of Excitons in Semiconductors 275
Fig. 13. Time dependences of the coherent exciton amplitude c0 and normal
(Z) and anomalous (£) self-energies after smooth switching on of a pump field
(at t = 0) for the pumping field strength / = 0.05 and pumping frequency
COQ = —1.05. Time is in units ofh/Ed.
of the equation
tanh z = (73)
1 - a z.
In such a case, two more branches of stationary states exist, not shown
in Fig. 11, with very large biexciton densities | d \>\ <70 |. It seems that
if a is not too close to 0.5, these densities appear to be too large for
excitonic molecules to really survive.
Fig. 14. The same as in Fig. 13, but / = 0.05, co0 = -1.025.
Fig. 15. The same as in Fig. 13, but / = 0.1, co0 = -1.05.
0 20 40 60 . 80 100 120 U0
Time
Fig. 16. The same as in Fig. 13, but / = 0.1, COQ = -1.025.
they even seem to become chaotic, which does not seem unexpected for
such a highly nonlinear system.
While small in reduced units, the field values supposed in Figs. 9-16
are really very large, corresponding to pumping intensities of the order
of tens and hundreds of GW/cm2. These phenomena can be observed
only in experiments with ultrashort pulses. Therefore, the details of the
relaxation mechanism become both important and interesting, as the
standard Boltzmann equation approach becomes inadequate. Different
theoretical approaches to this problem are presented in Refs. [71-73].
Being, in a sense, closely related to the problem of the coherence persis-
tance, it remains one of most important topics in the nonlinear optics of
semiconductors.
D.W. Snokef
Mechanics and Materials Technology Center
The Aerospace Corporation
Los Angeles, CA 90009-2957, USA
Abstract
Systematic studies of the photoluminescence of free excitons in CU2O re-
veal the temporal and spatial properties of these particles under a vari-
ety of conditions. Their relaxation and diffusion are measured by time-
resolved spectroscopy and imaging. At low densities the thermalization
and diffusivity are well characterized by deformation-potential theory- i.e.,
electron-phonon coupling. At high densities, the excitonic gas establishes
a well-defined temperature as a result of frequent interparticle collisions
and displays quantum statistics. The orthoexciton gas tends to saturate at
the phase boundary for Bose-Einstein condensation, characterized by the
ideal-gas relation n — CT 3^2. Under appropriate conditions, the paraex-
citon energy distribution exhibits an extra component which is interpreted
as a Bose-Einstein condensate.
281
282 J. P. Wolfe, J. L. Lin and D. W. Snoke
well-known wave function,
w = 1 e i (kr- W ( ) ; (1)
conduction band and a hole in the valence band bind together to form an
exciton with a center-of-mass wave vector kcm = ke + k/, and total mass
m = me + m^ The factor e2 in (5) and in ao is replaced by e2/e (where e
is the crystalline dielectric constant) and generally the mass of the hole is
comparable to that of the electron. In fact, the analogy between excitons
and positronium is even better than the analogy between excitons and
hydrogen, since the positron can be seen as a "hole" in the negative-
energy Dirac sea. The analogy between excitons and positronium is
better in CU2O than in most semiconductors, since the lowest conduction
band and the highest valence band are spherical and well separated
from other energy bands, and the effective electron and hole mass are
close to the free electron mass. This is illustrated by the fact that the
excitonic Rydberg in CU2O is almost exactly (13.6 eV)/2e2, the value
for positronium when e1 —• e2/e. Values of relevant parameters for the
crystal CU2O are summarized in Table 1. The Rydberg series of states
for excitons in CU2O has been clearly observed by optical absorption [1].
In the present review, we deal exclusively with the n = 1 ground state
because the thermal populations of the excited states are small, since the
binding energy of the exciton in CU2O is so large. The excitonic binding
energy Ex =150 meV is somewhat larger than the Rydberg, R^ = 91
meV, which describes the excited (n = 2,3,...) states, due to central-cell
corrections for this relatively small Wannier exciton.
Similar to the hydrogen atom, positronium and many atoms, the
exciton is a composite boson, having integer total spin. A collection
of excitons may be regarded as a gas of bosons provided the average
284 J. P. Wolfe, J. L. Lin and D. W. Snoke
distance between particles is much larger than the Bohr radius, or
-, (7)
or
na\ < -?-, (8)
for the interaction vertex (7(0) = 4naxh2/m. This condition is also fulfilled
for the experiments discussed here.
hv
(phonon assist)
h2k2
hv = £ g a p — Ex + (direct transition) , (9)
2m
h2k2
hv = £ Ex + 2m (phonon-assisted) , (10)
gap
where E gap is the semiconductor energy gap, Ex = e2/2ax is the ex-
citonic binding energy, | and hQt is the energy of an optical phonon
which is emitted or absorbed simultaneously with the photon emission.
Momentum conservation simultaneously requires for the two types of
processes,
k = kphoton (direct transition) (11)
T=74K
I There exists some confusion about whether electron-hole exchange gives a positive shift
to the ortho state, or a negative shift to the para state. In general, the answer depends
on the symmetry of the conduction and valence bands. For CU2O, the ortho state is
shifted up; as noted in Ref [6], the mass renormalization of the ortho (M ort h o = 3mo
instead of Mortho =me-\-m\x = 1.1 mo) implies a k-dependent positive exchange-energy
shift of the ortho state, analogous to that of the positronium triplet state.
288 J. P. Wolfe, J. L. Lin and D. W. Snoke
meV. As the temperature is lowered, the thermal population of optical
phonons decreases and the phonon-emission line dominates. The overall
temperature shift of the spectrum is due to a blue shift of the band gap,
£gap, as the crystal thermally contracts.
The dashed lines in Fig. 2 show an analysis of the low temperature
orthoexciton spectra. The phonon-assisted spectra are fit to a Maxwell-
Boltzmann distribution of the form,
where E is the kinetic energy of the exciton. The same fit temperature
T is assumed for the excitonic gas (giving the shape of the lines), and
for the crystal lattice (giving the relative intensity of the two lines,
/(emission)//(absorption) = ehQ/kBT). These data show unequivocally
that the orthoexcitons in weakly excited CU2O behave as a "classical
gas" in contact with a thermal reservoir, the lattice vibrations. The
lifetime of the orthoexcitons at low temperature is measured to be about
30 ns, so we conclude that the energy relaxation of these particles,
originating as hot electrons and holes, is sufficiently fast to reach thermal
equilibrium with the lattice within this time. As we shall see, the cooling
of the gas is accomplished by emission of phonons.
What is the cause of the 30 ns lifetime of the orthoexcitons? Recent
measurements indicate that the radiative lifetime of orthoexcitons is
much longer - several hundred nanoseconds [18]. It is most likely,
therefore, that the measured lifetime in this dilute-gas limit is due to
the decay of orthoexcitons to paraexcitons. Snoke et a\. [19] have
suggested that this interconversion process can take place by emission of
a single acoustic phonon. This mechanism appears to be consistent with
measurements of ortho-para interconversion times by Weiner et al. [20]
and with earlier measurements of luminescence intensities as a function
of temperature.
In contrast, the paraexcitons display lifetimes of many microseconds
in high quality naturally grown crystals [21], because their radiative
recombination is highly forbidden. | For a given experimental setup, the
3 Diffusion of Excitons
The long lifetimes of the paraexcitons have permitted Trauernicht et
al. [23] to observe their spatial diffusion by time-resolved luminescence
imaging. The two images of paraexciton luminescence in Fig. 3 show
the expansion of the paraexciton gas into the crystal. The diffusion
constant, D, can be determined by a straight-line fit to the square of the
cloud radius as a function of time [23]. A temperature dependence of
the measured diffusion constant is shown in Fig. 4. With each diffusion
measurement, the temperature of the paraexciton gas was obtained by
fitting its spectrum to the Maxwell-Boltzmann distribution (13).
The measured diffusion constants are huge, ranging up to 1000 cm 2 /s
for the lowest experimental temperature of 1.2 K. The strong temperature
dependence indicates that this process is mediated by thermal vibrations
of the lattice, i.e., exciton-phonon scattering. Taking into account the
Planckian distribution of phonons at temperature T, and making the
high-temperature approximation k%T > mv2L, where VL is the lattice
t Broadening of the spectrum associated with the lifetime of the excitons is negligible.
The Lorentzian broadening from the uncertainty principle implies a spectral full width
at half-maximum of 1.2 meV for a lifetime of 1 ps. Therefore, the broadening of the
orthoexciton spectrum due to a 30 ns lifetime is an unmeasureable 0.04 //eV. Also,
broadening due to exciton-phonon interactions, with a typical scattering time of 30 ps,
leads to a linewidth of only 0.04 meV at low temperature. On the other hand, interparticle
collisions can cause a measurable broadening at high gas densities [13].
290 J. P. Wolfe, J. L. Lin and D. W. Snoke
-crystal surface
Fig. 3. Time-resolved images of the paraexciton luminescence of Cu2O at T=2K,
obtained by x-y scanning of the crystal image across an entrance aperture of
a spectrometer, (a) £=0.2 fis after a 100-ns Ar + laser pulse is absorbed at the
crystal surface, (b) £=0.6 /is.(From [23].)
1000
CM
100
-r-l/2
10 10 30 100
T (K)
Notice that in (16), the scattering rate varies inversely as the square
of the phonon velocity. For spherical electron and hole bands (as in
CU2O), only longitudinally polarized phonons can couple to the carriers.
Trauernicht et al. [24] show that if the crystal is stressed, the reduced
symmetry allows the excitons to couple also to transverse phonons, which
have a velocity about four times smaller than that of the longitudinal
phonons. Hence, an externally applied stress acts to reduce the diffusivity
of the excitons and may be an effective means of confining the excitons
to a smaller volume.
time is T = l/vav « 20 ns, which implies a mobility of x/m = 10 8 cm 2 /eV-s, perhaps the
highest measured for any electronic species in a non-superconducting bulk crystal.
292 J. P. Wolfe, J. L. Lin and D. W. Snoke
to 1 — /k, where /k is the occupation number of the state. This is the
Pauli exclusion principle at work.
The exciton, being composed of two spin-1/2 particles, is an integral-
spin particle, or composite boson. The quantum statistical rule for a
boson is that its probability of being scattered into the state of wavevector
k is proportional to 1 + /k- In other words, the likelihood that a boson
is scattered into a given state is enhanced if other bosons already occupy
that state. It can be shown [25] that this "stimulated" scattering leads to
the Bose-Einstein distribution function,
For a given density and temperature, (21) fixes the chemical potential,
fi, of the gas. By defining e = E/kBT and the "dimensionless chem-
ical potential" a = —fi/k BT, Eq. (21) can be written in terms of a
dimensionless integral as,
e l/2
-^-jde. ill)
BEC of a Nearly Ideal Gas: Excitons in 293
e = E/kT
Fig. 5. The Bose-Einstein distribution N(E) = f{E)D(E) for several values of a
for an ideal three-dimensional gas at constant potential.
f Fermi broadening of degenerate electron-hole plasmas has been clearly observed, for
example, in electron-hole droplets in Ge and Si [27].
294 J. P. Wolfe, J. L. Lin and D. W. Snoke
to obtain
100
BEC phase boundaries
//Vff,
m = 2.7m 0 / /
fc 10 - III!, / /
/ cond ensate /
10" 'ifttl/ll/,7/
IO IC
' / / / / , / /
I019
Density (cm 3 )
1020
— = nvavG, (26)
0 10 20 30 40
Kinetic energy (meV) \J 5 10
3 IU 15
ID
Kinetic energy (meV)
Fig. 7. (a) Theoretical particle distribution for a boson gas undergoing only
interparticle scattering, for n « nQ. The label of the curves is the number of
scattering events per particle since the initial creation. The last curve is within
1% of a Maxwell-Boltzmann distribution. (From Ref. [35].) (b) Theoretical
particle distribution for a boson gas undergoing only inlelastic scattering with
phonons, assuming an unchanging phonon bath temperature below the critical
temperature for condensation of the gas. The particle mass and particle-phonon
scattering matrix element are those of excitons in CU2O, but the model does not
take into account exciton-exciton interactions important in this density regime.
(From Ref. [6].)
t(ps)
0.0
10.8 —
c
c 21.6
CD
32.4 E
U
N
43.2
W
E 10
54.0
64.8
Fig. 8. (a) Time-resolved luminescence from CU2O following a very short (5 ps)
laser pulse at 2.033 eV, measured with a streak camera. The dashed lines are
a fit to the solution of the Boltzmann equation for exciton-phonon scattering
based on deformation-potential theory. (From Ref. [6].) (b) The energy loss rate
for excitons in a Maxwell-Boltzmann distribution and zero phonon temperature,
using the phonon emission rates deduced in Ref. [6].
hQt < 20 meV. The most favored process is emission of the 18.7-meV
TO phonon, apparent in the data of Fig. 8(a) in the peak near zero
kinetic energy. The other sharp peaks are longitudinal-acoustic (LA)
phonon emission peaks. Phonon emission depletes the 20-meV kinetic
energy states in about 30 ps, rather slow compared to semiconductors
such as GaAs, but still much less than the exciton lifetime. At 64.8 ps,
the distribution resembles a Maxwell-Boltzmann distribution with T «
50 K, which is still considerably higher than the 18 K lattice temperature.
The exciton gas does not reach the lattice temperature until hundreds of
picoseconds later (see Fig. 12). Experiments such as these give the matrix
elements for the optical and acoustic phonon emission processes. Fig. 8(b)
is a summary of the energy loss rate for an exciton gas at temperature
T, with the lattice temperature assumed to be zero. The dotted curve is
300 J. P. Wolfe, J. L. Lin and D. W. Snoke
the T3/2 behavior predicted in Section 3 for acoustic-phonon emission.
Deviation from this line at lower energies is due to the "freeze-out"
process indicated in Fig. 4. The enhancement at higher temperatures
is due to optical-phonon emission. The total of these processes gives a
temperature dependence of the rate of energy loss to phonons of roughly
T5'2 in the range 1-100 K.
Finally, we can ask under what conditions the ortho and para compo-
nents of the exciton gas reach chemical equilibrium. As mentioned above,
orthoexcitons can decay via a phonon-assisted process with a time scale
of 30 ns at 2 K and about 600 ps at 30 K [19, 20]. f Phonon-assisted
paraexciton conversion to the orthoexciton state lying 12 meV higher
requires phonon absorption, which is highly suppressed at low tempera-
ture. At high densities, however, the two components can "convert" into
each other via the Auger process discussed above, which destroys one
exciton and ionizes the other. The free carriers thus created can form
into either kind of exciton. This leads to a relatively efficient para-to-
ortho "up conversion" - orthoexcitons have been observed even at 2 K
at moderate densities [37]. Because this process is so fast, ortho and
para excitons will in general never be in chemical equilibrium, i.e. have
population ratio N0/Np = 3exp(—A/k BT), at densities above 1017 cm"3
and lattice temperatures below 30 K; instead, their relative number will
be determined by a rate balance between conversion processes. The two
components will have the same temperature but two different "effective
chemical potentials", oc0 and ocp.
If the two components were in chemical equilibrium, the orthoexcitons
would hardly ever have on0 less than A//CB T « 4 at 30 K, and they would
never deviate from a Maxwell-Boltzmann distribution (see Fig. 5) at low
temperature. As we shall see, in fact the orthoexcitons exhibit quantum
statistics over a wide range of density and temperature. The ortho-to-
para conversion process can be enhanced, however, by raising the lattice
temperature, pushing the gas toward chemical equilibrium.
Table 3 gives a summary of the various rates of relaxation processes
for excitons in CU2O. Relaxation and recombination processes of excitons
occur in CU2O on time scales which vary from picoseconds to millisec-
onds, and properly modeling this system requires understanding which
processes are important in a given density and temperature regime.
t=16ns $ W K ^ (a)
c
\
/
CD ~\
CD
N
"5
\a V
E
|V"T(OO
t=44 ns | h
r (b)
c
CD
CD
o
E T=32K ^
/
o
-4 0 4 8 12 16 20 24 28 32 36 40 44 48
t(ns)
Fig. 9. (a) The orthoexciton (dashed line) and paraexciton (solid line) energy
distributions during laser-pulse excitation at photon energy of 2.4 eV. The two
spectra are normalized to the same height and shifted in energy for comparison.
The orthoexciton distribution fits a Maxwell-Boltzmann with T=38 K. The
paraexciton spectrum is obscured by orthoexciton luminescence at energies above
2.015 eV. The open circles are a 1.3 meV-broadened Bose-Einstein distribution
with a = 0.07 and T=38 K. (b) The distributions well after the laser pulse.
Both fit a Maxwell-Boltzmann distribution with T=32 K (open circles), (c) The
spectral FWHM of each energy distribution as a function of time (open circles:
orthoexcitons; solid circles: paraexcitons). The 10-ns laser pulse is centered at
£=10 ns. (From Ref. [39].)
304 J. P. Wolfe, J. L. Lin and D. W. Snoke
(c)
Fig. 10. (a) Excited-states Bose-Einstein distribution for a gas at constant po-
tential with a=0. (b) The distribution of (a) convolved with a triangular spectral-
resolution function, (c) The distribution of (a), plus a delta-function peak at
E = 0. (d) The distribution in (c), convolved with the same spectral resolution
function.
t-i20ps (a)
unjts)
t«900 ps
•e ^ a—0.1
o
1
~ 10
;
1000
fi
100-• o High density
°O
0
o O O
<» ° o o o o o
10-- Kl^ T
Low density -—.
T,ott
1 1 1
200 400 600 800 1000
t(ps)
Fig. 12. The effective temperature as a function of time for the case of high
density (n > 1019 cm" 3 ) and low density (n < 1017 cm" 3 ). The solid line in the
low density case is the prediction of the deformation-potential theory for phonon
emission discussed in the text. (From Ref. [38].)
the gas is evolving along a line parallel to the (/i = 0) phase boundary
for BEC, as shown by the open circles in Fig. 14. The effect of the gas
to remain close to the phase boundary over a wide range of densities
and temperatures has been termed "quantum saturation" by Snoke et
al. [42]. This remains one of the major puzzles of this remarkable
quantum gas.
Quantum saturation is even more pronounced in a "long-pulse" (10 ns)
experiment. In this case, the orthoexciton gas is observed with about 1 ns
resolution during and after the excitation pulse. Again, the gas tempera-
ture and chemical potential are found to be well defined throughout the
experiment. As the excitation pulse evolves, the temperature and density
of the gas increase along a \i « —O.l/c^T line to maximum values (at
about the peak of the pulse) and then decay along the same line, as shown
by the solid dots in Fig. 14. In these experiments, the signal levels are
high enough to observe the phonon-assisted spectra of the paraexcitons
simultaneously. At medium pulse intensities, the paraexcitons exhibit a
higher degree of degeneracy, as predicted in Fig. 6. At high pulse inten-
sities, interesting anomalies appear in the paraexciton spectrum which
were attributed to the formation of a condensate.
Clues to the origin of the quantum saturation effect may be extracted
from the observations of an increased orthoexciton decay rate and a
reduced cooling rate (Figs. 12 and 13). Figure 15 explicitly plots the
308 J. P. Wolfe, J. L. Lin and D. W. Snoke
I
•"fit (a)
1019 • s -
Eo
l0 18< <t • •
Q)
Q 1 •Laser intensify
•
m17 (
0.25
0.00
(b)
-0.25 II
-0.50
-0.75
-1.00
200 400 600 800 1000 1200
t(ps)
Fig. 13. The thermodynamic parameters of the orthoexciton gas in the case
shown in Fig. 11. At late times the uncertainty in the chemical potential \i grows
as the temperature lowers (Fig. 12) and the line narrows. The smallest possible
a is plotted as a solid dot, but, within the spectral resolution, the corresponding
spectra at late times are almost compatable with a classical distribution, as
indicated by the large error bars. (From Ref. [38].)
100 •
1017 1 0 18 10 19 1020
Density (cm" 3 )
Fig. 14. The quantum saturation of the orthoexciton gas, corresponding to a
temperature rise with increasing density, T oc n2/3. Solid circles are deduced
from fits to spectra during and after long (10-ns) laser pulses. (From Ref. [42].)
Open circles are deduced from fits to spectra following short (100-ps) laser pulses.
(From Ref. [38].) The dashed line gives the critical densities for BEC of the g = 3
orthoexcitons.
1018 1019
Density (cm"
Fig. 15. The decay rate of orthoexcitons as a function of density, from the same
data as Fig. 12. (From Ref. [38].)
€2
g=3
2500
t = 6.4 ns
2000 cr =1.0 k b a r
T=40 K
(fit) -
-500
2020 2021 2022 2023 2024 2025
Photon energy (meV)-ortho
Fig. 16. Orthoexciton luminescence from Cu2O under [110] stress.
(a) t=9.6 ns
(b)
,6 2000 ft cr=1.0 kbar
jp
;cence intensity
1500
-
1000
/ - expt.
0 fit (T=62K, a ^0.015)
VI
<D
500
L MB V - MB (T=62K)
e 0 J 1 1 10" 1 ,
2020 2021 2022 2023 2024 2025 2020 2021 2022 2023 2024 2025
Photon energy (meV)
Fig. 17. Solid line: orthoexdton energy distribution in the lowest spin state in
stressed CU2O following intense laser exdtation. (Luminescence from higher spin
states has been subtracted via a polarizer.) Open circles: fit to the ideal-gas
Bose-Einstein distribution, (a) Linear plot, (b) Semilog plot.
100
# stressed
O unstressed
BEC boundary
10 10 •
associated with the degenerate quantum nature of the gas and is not
solely due to an "accidental" coincidence of classical heating processes
(e.g., Auger or ortho-para conversion) with the BEC phase boundary.
316 J. P. Wolfe, J. L. Lin and D. W. Snoke
200
3 A A K
Fig. 19. The orthoexciton (dashed line) and paraexciton (solid line) energy dis-
tributions during laser-pulse excitation at photon energy of 2.4 eV in a [110]
stressed crystal. The two spectra are normalized to the same height and shifted
in energy for comparison.
a=0.3 kbar
• i
' • • • •
10 J
: / \ para (x500)
• c 0 ortho
O -
O
O
S7 oo(D O
_v v O
O
V V
laser V
V
o
V
o
7 . I . I .
10 20 30 40 50 60
Time (ns)
Fig. 20. The relative intensity of the ortho and para luminescence. Triangles give
the intensity of the laser vs. time, multiplied by an arbitrary constant.
(b)
Fig. 21. (a) Time evolution of the temperature and chemical potential determined
from the orthoexciton spectrum during and after an intense 10-ns laser pulse of
a Cu2O crystal stressed along [110] at 0.3 kbar. (b) Density and temperature for
the ortho and para gases during the same experiment. Labels for the para data
points indicate the elapsed nanoseconds.
f That the orthoexciton and paraexciton gas occupy nearly the same volume during
the (10 ns) excitation pulse is quite reasonable considering that there are frequent
interparticle collisions to keep them together at high densities, and that the orthoexcitons
convert to form a good fraction of the paraexcitons. An imaging experiment with high
spatial resolution in an unstressed crystal has shown that the ortho and paraexciton
gases coexist spatially at least up to t = 25 ns (see Ref. [13].)
BEC of a Nearly Ideal Gas: Excitons in CujO 319
Photon e n e r g y r (meV)-ortho
2021.2 2022.2 2023.2
1
i i
t=15 ns
h
600 CJ^O.S kbar
ortho
para
O fit
400 -
\ • (T=19K,a p =0)
S
^N - (A=0.5 meV)
A"
200 -
x
A k K C p ""
p
•
0(
Fig. 22. The data of Fig. 19, fit to the BEC distribution discussed in the text.
Nc = 1202g(kBT/ha))\ (31)
For a given temperature, when the total number of particles in the well
exceeds iVc, the excess particles will theoretically occupy the ground state.
Herein lies the qualitative departure from the uniform-potential case:
since the ground state of the harmonic oscillator is a narrow Gaussian
wavefunction centered at the bottom of the well, BEC corresponds to a
spatial condensation rather than a fc-space condensation.
The spatial extent of the ground state for an ideal gas, calculated
simply as the wavefunction of a single particle in the ground state of
the well, is so small that the condition (6) for the excitons to remain
bosons, na3 <C 1, would not hold if a significant fraction of the gas
were condensed in this state. In this case, one cannot ignore the effect of
322 J. P. Wolfe, J. L. Lin and D. W. Snoke
Laser
(a) (b)
Fig. 23. (a) Equipotential curves for a Hertzian contact stress, (b) Drift of excitons
into a three-dimensional parabolic well created in a CU2O crystal with Hertzian
contact stress. This photoluminescence image of paraexcitons is obtained for
continuous excitation at the left surface with an Ar + laser. (From Ref. [54].)
o
15
<7>
Fig. 24. Spatial profiles of the photoluminescence from paraexcitons in the po-
tential well, during cw excitation with a dye laser. Here r is the distance from
the center of the well. (From Ref. [54].)
i(D
Q
n (2. \V1
(33)
The condensation condition for the two cases can be compared by noting that nc
BEC of a Nearly Ideal Gas: Excitons in Cu2O 325
for Auger-limited lifetimes, where
12 Conclusions
Despite their short lifetimes, free excitons in a high-purity crystal of
CU2O form a nearly ideal gas of particles. The excitonic gas obeys the
kinetic theory of gases, it moves via drift or diffusion through the crystal
long after the generating laser is off, it can be trapped in a potential
minimum, and on timescales longer than a nanosecond it exhibits a
Maxwell-Boltzmann or Bose-Einstein energy distribution characteristic
of an internal thermal equilibrium. Because the exciton lifetime is much
longer than the particle scattering time, the laws of equilibrium thermo-
dynamics apply. Indeed, the finite lifetime of the excitons has a distinct
benefit: due to the radiative decay of excitons, it is possible to observe
their energy distribution directly.
The search for BEC of excitons has progressed stepwise. First, it has
been proven that excitons behave statistically as bosons- namely, the
energy distribution narrows with increasing density, in contrast to the
broadening of the fermion distribution with increasing density. Second,
the BEC phase boundary has been mapped over more than an order
of magnitude of density variation, via the "quantum saturation" effect.
Third, it has been shown that the density of excitons can exceed the
critical density for condensation at a given temperature, under certain
conditions. Finally, an extremely narrow energy distribution (A£ <C fc^T)
of paraexcitons in the ground state has been observed under conditions
of n > nc.
A full understanding of these results awaits a complete theoretical
treatment of the weakly interacting exciton gas, including the hydro-
326 J. P. Wolfe, J. L. Lin and D. W. Snoke
dynamics of the expanding boson gas and the energy distribution of a
finite-lifetime, two-component gas with phonon and Auger interactions.
Perhaps the most significant lack in present theory is an exact calculation
or an accurate measurement of the exciton-exciton interaction potential:
the proper form can not be simply extrapolated from that of hydrogen,
due to the important effects of the light mass of the hole and the electron-
hole exchange. This interaction potential goes into the calculation of the
biexciton binding energy, the ortho-para splitting and interconversion,
and exciton-exciton scattering.
Nevertheless one of the beauties of this system is the fact that the laws
for an ideal Bose gas work so well over such a large range of density and
temperature, attested by the fact that the spatial diffusion data and the
spectra are well fit using simple Maxwell-Boltzmann or Bose-Einstein
distributions. All of this analysis supports the conclusion at this point
that paraexcitons do indeed undergo Bose-Einstein condensation.
References
[I] S. Nikitine, J.B. Grun, and M. Sieskind, J. Phys. Chem. Solids 17,
292 (1961).
[2] M. O'Keefe, J. Chem. Phys. 39, 1789 (1963).
[3] A. Gotzene and C. Schwab, Solid State Comm. 18, 1565 (1976).
[4] J.W. Hodby et al, J. Phys. C 9, 1429 (1976).
[5] P.Y. Yu and Y.R. Shen, Phys. Rev. Lett. 32, 939 (1974); Phys. Rev.
B 12, 1377 (1975).
[6] D.W. Snoke, D. Braun, and M. Cardona, Phys. Rev. B 44, 2991
(1991).
[7] V.T. Agekyan, Phys. Stat. Sol. (a) 43, 11 (1977).
[8] RD. Bloch and C. Schwab, Phys. Rev. Lett. 41, 514 (1978).
[9] E.g. A.L. Fetter and J.D. Walecka, Quantum Theory of Many-Particle
Systems (McGraw-Hill, New York, 1971), p. 259.
[10] PL. Gourley and J.P. Wolfe, Phys. Rev. B 25, 6338 (1982).
[II] M.M. Beg and S.M. Shapiro, Phys. Rev. B 13, 1728 (1976).
[12] Y. Petroff, P.Y. Yu, Y.R. Shen, Phys. Rev. Lett. 29, 1558 (1972);
Phys. Rev. B 12, 2488 (1975).
BEC of a Nearly Ideal Gas: Excitons in Cu2O 327
[13] D.W. Snoke, J.R Wolfe and A. Mysyrowicz, Phys. Rev. B 41 11171
(1990).
[14] K. Reimann and K. Syassen, Phys. Rev. B 39, 11113 (1989).
[15] The measured relative absorption ratio of 50 in Ref. [8] is reduced
by a factor of three to take into account the orthoexciton spin
multiplicity.
[16] J.L. Birman, Sol. State Comm. 13, 1189 (1978).
[17] F. Bassani and M. Rovere, Solid State Comm. 19, 887 (1976).
[18] D.W. Snoke, A J . Shields, and M. Cardona, Phys. Rev. B 45, 11693
(1992).
[19] D.W. Snoke, J.R Wolfe, and D.P. Trauernicht, Phys. Rev. B 41, 5266
(1990).
[20] J.S. Weiner et al, Solid State Comm. 46, 105 (1983).
[21] A. Mysyrowicz, D. Hulin, and A. Antonetti, Phys. Rev. Lett. 43,
1123 (1979).
[22] K. Reimann, J. Lum. 40/41, 475 (1988); A.V. Akimov and A.A.
Kaplyanskii, Proc. 5th Int. Conf. on Phonon Scattering, A.C. Ander-
son and J.R Wolfe, eds. (Springer, Berlin, 1986), p. 449.
[23] D.P. Trauernicht, J.R Wolfe, and A. Mysyrowicz, Phys. Rev. Lett.
52, 855 (1984).
[24] D.P. Trauernicht and J.P. Wolfe, Phys. Rev. B 33, 8506 (1986).
[25] E.g. D.I. Blokhintsev, Quantum Mechanics (Dordrecht-Holland, Am-
sterdam, 1964), p. 493ff.
[26] E.g. F.K. Richtmyer, E.H. Kennard, and John N. Cooper, Introduc-
tion to Modern Physics, 6th edition (McGraw-Hill, New York, 1969)
p. 571.
[27] For a review see J.C. Hensel, T.G. Phillips, and G.A. Thomas, Solid
State Physics 32, H. Ehrenreich, F. Seitz, and D. Turnbull, eds.
(Academic, New York, 1977).
[28] T.J. Greytak and D. Kleppner, in New Trends in Atomic Physics, G.
Grynberg and R. Stora, eds., (North-Holland, Amsterdam, 1984).
[29] R.C. Casella, J. Phys. Chem. Solids 24, 19 (1963).
[30] A. Mysyrowicz, D. Hulin, and E. Hanamura, in Ultrafast Phenomena
VII, (Springer, Berlin, 1990).
[31] H. Haug and H.H. Kranz, Z. Phys. B 53, 153 (1983).
[32] A. Quattropani and J.J. Forney, II Nuovo Cimento 39, 569 (1977).
[33] A.I. Bobrysheva and S.A. Mosklenko, Phys. Stat. Sol. (b) 119, 141
(1983); S.A. Moskalenko et al, J. Phys. C 18, 989 (1985); Phys. Stat.
Sol. (b) 129, 657 (1985); A.I. Bobtysheva et al, Phys. Stat. Sol. (b),
167, 625 (1991).
328 J. P. Wolfe, J. L. Lin and D. W. Snoke
[34] D.W. Snoke and J.R Wolfe, Phys. Rev. B 39, 4030 (1988).
[35] D.W. Snoke, W.W. Riihle, Y.-C. Lu, and E. Bauser, Phys. Rev. B 45,
10979 (1992).
[36] Semiconductors Probed by Ultrafast Laser Spectroscopy (Academic
Press, New York, 1984).
[37] D.R Trauernicht, A. Mysyrowicz, and J.R Wolfe, Phys. Rev. B 28,
3590 (1983); D.R Trauernicht, J.R Wolfe, and A. Mysyrowocz, Phys.
Rev. B 34, 2561 (1986).
[38] D.W. Snoke and J.R Wolfe, Phys. Rev. B 42, 7876 (1990).
[39] D.W. Snoke, J.-L. Lin, and J.R Wolfe, Phys. Rev. B 43, 1226 (1991).
[40] Bose narrowing was shown for excitons in stressed Ge by V.B. Tim-
ofeev, V.D. Kulakovskii, and I.V. Kukushkin, Physica B+C 117/118,
327 (1983).
[41] D. Hulin, A. Mysyrowicz, and C. Benoit a la Guillaume, Phys. Rev.
Lett. 45, 1970 (1980).
[42] D.W. Snoke, J.R Wolfe, and A. Mysyrowicz, Phys. Rev. Lett. 59, 827
(1987).
[43] A. Mysyrowicz, D. Hulin, and C. Benoit a la Guillaume, J. Lum.
24/25, 629 (1981).
[44] L.V. Gregor, J. Phys. Chem. 66, 1645 (1962).
[45] D.W. Snoke, University of Illinois at Urbana-Champaign Ph.D.
thesis, 1990.
[46] For a hydrodynamic model of excitons in CU2O without Auger
recombination and phonon emission, see B. Link and G. Baym,
Phys. Rev. Lett. 69, 2959 (1992).
[47] See the contribution of Yu. Kagan in this volume.
[48] H. Shi, G. Verechaka and A. Griffin, Phys. Rev. B 50, 1119 (1994)
have derived an expression for the decay luminescence spectrum
for an excitonic BEC in the weakly interacting gas approximation,
incorporating the low-energy tail and the blue shift of the spectrum.
[49] J.-L. Lin and J.R Wolfe, Phys. Rev. Lett. 71, 1223 (1993).
[50] R.G. Waters et a/, Phys. Rev. B 21, 1665 (1980).
[51] H.-R. Trebin, H.Z. Cummins, and J.L. Birman, Phys. Rev. B 23, 597
(1981).
[52] A. Mysyrowicz, D.R Trauernicht, J.R Wolfe, and H.-R. Trebin, Phys.
Rev. B 27, 2562 (1983).
[53] S.A. Moskalenko et a/., Phys. Stat. Sol. (b) 129, 657 (1985).
[54] See D.R Trauernicht, J.R Wolfe and A. Mysyrowicz, Phys. Rev. B
34, 2561 (1986).
BEC of a Nearly Ideal Gas: Excitons in CU2O 329
[55] T.W. Hijmans, Yu. Kagan, G.V. Shlyapnikov, and J.T.M. Walraven,
to be published; see also D.A. Huse and E.D. Siggia, J. Low Temp.
Phys. 46, 137 (1982) and V.V. Goldman, I. Silvera, and AJ. Leggett,
Phys Rev. B 24, 2870 (1981).
[56] J.P. Wolfe and C D . Jeffries, in Electron-Hole Droplets in Semi-
conductors, C D . Jeffries and L.V. Keldysh, eds. (North-Holland,
Amsterdam, 1987).
14
Bose-Einstein Condensation of Excitonic
Particles in Semiconductors
A. Mysyrowicz
LOA-ENSTA
Ecole Poly technique
F-91120 Palaiseau
France
Abstract
The case of excitons as candidates for Bose-Einstein condensation is dis-
cussed, and experimental results in CuCl and CU2O are presented. In CuCl,
spectral analysis of the luminescence from biexcitons as a function of their
density reveals a gradual evolution from classical statistics towards a quan-
tum degenerate regime. The appearance of a sharp emission line below a
critical temperature and above a critical density is attributed to the pres-
ence of a laser-induced Bose-Einstein condensate of excitonic molecules.
This interpretation is supported by pump-probe experiments which show
that additional particles injected in the presence of a biexciton condensate
are drawn into it.
In Cu^O, free exciton luminescence spectral analysis of ortho- and
paraexcitons reveals a gradual evolution from a classical to a Bose quantum
degenerate regime with increasing particle densities. Orthoexciton densities
close to the critical density for condensation are obtained at high inco-
herent excitation. Under similar pumping, paraexciton densities exceeding
the critical value are inferred from luminescence intensity ratios. Anoma-
lous transport properties of paraexcitons, such as ballistic propagation over
macroscopic distances and formation of soliton-like excitonic packets are
discussed as evidence for excitonic superfluidity.
1 Introduction
1.1 Excitons
The lowest electronically excited state of a non-metallic crystal corre-
sponds to the promotion of one electron from the top of the highest fully
330
BEC of Excitonic Particles in Semiconductors 331
occupied valence band to the bottom of the next empty conduction band.
A correct evaluation of the required energy must include the Coulomb
correlation between the promoted electron and all other electrons left
behind in the valence band. A very fruitful conceptual approach (see the
article by Wolfe et al. in this volume) consists in treating this problem
by the equivalent point of view of the Coulomb interaction between a
negatively charged particle, the electron in the conduction band and a
positively charged particle, the electron vacancy or positively charged
hole left in the valence band. The problem reduces to the same two-
particle Schrodinger equation as for the hydrogen or positronium atom,
h2 2 h2 e2 1
Ve V
~2m~r ~ ^~ * ~ ~ ~ I °= E®' W
It is well-known that discrete solutions exist corresponding to bound
electron-hole pair states. These bound states are called excitons and
obey the hydrogenic relation
1.2 Biexcitons
It is possible to pursue the analogy between excitons and positronium or
hydrogen atoms one step further and consider the chemistry occurring at
higher densities when excitons start to interact. Like hydrogen atoms, ex-
citons can form molecules called biexcitons. The formation of a biexciton
due to the coupling of a pair of excitons and its analogy to a hydrogen
molecule has been experimentally well established. An excitonic molecule
in its ground state is similar to a hydrogen paramolecule with two holes
of opposite spins sharing two electrons, also with opposite spins. It can
BEC of Excitonic Particles in Semiconductors 333
also propagate through the sample, carrying a total energy:
where 2E^ = 2(Eg — Bx) is the internal energy of two noninteracting exci-
tons at zero wave vectors, E®x is the biexciton energy at zero wave vector,
Dxx is the biexciton binding energy (the energy required to dissociate it
into two free excitons), and Kxx and Mxx are the center-of-mass wave
vector and mass of the biexciton, respectively.
The binding energy of the biexciton depends on the ratio of effective
masses, a = me/mh [2]. The limiting case a = 0 corresponds to the hydro-
gen molecule, with a well-established experimental binding energy. No
experimental verification of the other limit, a = 1 (molecule of positro-
nium) exists yet. However, variational calculations predict a biexciton
binding energy smaller by one order of magnitude.
Due to the small mass of excitons, typically of the order of the free
electron or less, quantum fluctuations are very important in determining
the internal structure of biexcitons. The biexciton zero-point energy is
comparable to the molecular binding energy itself. As a consequence,
vibrational levels are not stable and at most one excited level exists in
crystals with simple band structures (the orthobiexciton in the first rota-
tional level). This is in sharp contrast to the rich manifold of rotational
and vibrational levels of hydrogen molecules. This simplicity of the
molecular excitation spectrum is favorable for Bose condensation since
it prevents formation of a molecular liquid phase, as will be discussed
shortly.
Biexcitons cannot be created by optical means directly into their ground
state through a one-photon absorption process, since the required photon
energy is almost twice the band gap energy and therefore corresponds
to a region of strong absorption in the exciton ionization continuum.
It is possible to create biexcitons directly from the crystal ground state
by two-photon absorption. As shown by Hanamura [3], the two-photon
absorption cross section corresponding to the direct creation of a para-
biexciton has giant values, of the order /? « 1 cm- MW" 1 , or 5-6 orders
of magnitude more than for typical two-photon interband transitions in
crystals. Such huge values reflect the fact that a two-photon process
is well adapted to the creation of a biexciton, because two electrons
separated by a mean distance equal to the biexciton radius must be
simultaneously promoted from the valence to the conduction band. Con-
sequently, densities of biexcitons reaching 1019 cm"3 can be obtained
334 A. Mysyrowicz
with a laser of intensity less than 1 MW/cm2 [4]. In the context of
Bose condensation, it is important to note that the biexciton population
is nearly spatially uniform, since the two-photon attenuation of a laser
decreases as a function of crystal thickness z as
<4>
instead of the usual exponential attenuation from Beer-Lambert law
in a one-photon process. If two independent laser beams are used
to generate biexcitons directly, the wavevector of the created particles
can be controlled between 2Ko and 0 by changing the angle between
the incoming beam, where Ko is the magnitude of the incident photon
wavevector.
Biexcitons can also be created indirectly through free carrier (or ex-
citon) generation followed by biexciton formation [5]. In a steady state
regime, the ratio of biexciton to exciton populations Nxx/Nx is dictated
by the values of their respective chemical potentials. Biexciton densities
reaching 1020 cm"3 have been obtained through indirect band-to-band
pumping [6]. Note however that under indirect creation process, the
biexciton density is spatially nonuniform since it is proportional to the
local exciton density, itself spatially nonuniform. Also, the excess pump
photon energy leads to local heating of the exciton and biexciton gas.
where
(6)
: 2M?
and
(7)
where q is the emitted photon wave vector.
The biexciton luminescence lineshape can be expressed as
fi
where the parameter ft characterizes the oscillator strength of the ex-
citonic transition. The longitudinal and uncoupled transverse exciton
energies are
ET(K)=ET(0)
-exciton
EM/2
K
h2c2K2 4TT p
= 1 + 1-E2/E2T(KY (12)
I i 1
CuCI Excitation for K = 0
TL = 25K
^t>^V y >o |I = - oo
J
tensi
.g
TL = 5K
O
<D
o IT H
</> XT ^a P A. T =25
umi ne
p d ««
/
I I i
3.170 3.165 3.160
Photon energy (eV)
Fig. 2. Luminescence spectrum of CuCI under weak (upper) and strong two-
photon excitation. The open circles are calculated lineshapes assuming a classical
and a Bose quantum distribution with n — 0.
for the particle distribution, indicating that the critical density for BEC
is reached. Additional measurements for intermediate pump intensities
show a gradual evolution between these two limiting cases. Also, mea-
surements as a function of sample temperature show a disappearance of
quantum degeneracy in the biexciton population above T « 40 K, irre-
spective of the pump intensity, in good agreement with the predictions
from the ideal Bose gas [19].
BEC of Excitonic Particles in Semiconductors 341
Fig. 3. Biexciton luminescence with single beam excitation for several temper-
atures of the sample. The spectra are shifted so that the sharp line NT lines
coincide. The shift is due to the change of band gap energy with T.
Above a critical input intensity a sharp line, denoted NT, grows on the
high energy side of the Mr band if a single-beam, resonant two-photon
excitation is used. This line shows remarkable properties, such as a large
spatial anisotropy, even though the crystal has cubic symmetry. It is
intense in the backward detection geometry (with respect to the incident
beam), but very weak in the forward geometry [19]. Further, the strength
of the line shows a marked dependence upon sample temperature (see
Fig. 3), with a threshold temperature for appearance varying with input
excitation in the manner shown in Fig. 4. The appearance of iVj has been
interpreted as evidence for the occurrence of Bose-Einstein condensation
in the biexciton gas [19], with the condensation taking place at the
wavevector imposed by the incident excitation. Since the radiative decay
342 A. Mysyrowicz
CuCl
2.8 |im film
20K
i
%
8
£
0.025 In(x1500)
1 1 i i i i
k ^J
| \
\
\
X
probe
pump only
Xi
J
Uo/2
M V
\
>v
pump only
x2
^^(no pump)
I \ probe
\
^^^ ^ _ (witn pump )
j ^ \ probe
V (with pump)
^S. x32
i i
3.175 3.170 3.165 3.175 3.170 3.165
Photon energy (eV)
into a reflected wave Er(r,t) with the temporal oscillation unchanged but
a reversed spatial phase
= E(r,-t) (16)
Therefore, the reflected wave Er(r, t) behaves as if it retraces the path
of the incoming wave back in time, as in a movie played backwards.
This intriguing property leads to automatic compensation, after reflec-
tion, of distortions of the phase front experienced by an incident wave
on its way to the conjugate mirror. It is possible to take advantage
of this property to perform spectroscopy in the space domain instead
of the frequency domain. With a small aperture one can discriminate
a coherent wave of well-defined K-vector reflected by conjugate mir-
ror from other radiations of the same frequency but different K -vectors
originating from the same mirror (see Fig. 6). A true phase conjugate
mirror is obtained by inducing a coherent, spatially uniform oscillation
at frequency co. The medium acts as a two-photon amplifier, ampli-
fying an incident beam (co,Ki). In nonlinear optics, the macroscopic
oscillation at 2co is induced by two counterpropagating beams of the
BEC of Excitonic Particles in Semiconductors 345
50% mirror
to detector | /
conjugate mirror
pinhole "™ ™"
i i
D = 600 cm2/sec
0 20 40 60 80 100
t(ns)
Fig. 8. Expansion of the orthoexciton cloud towards the interior of the crystal,
as measured by time- and space-resolved luminescence, for three different initial
exciton densities, where A is the distance of half maximum of the luminescence
intensity from the surface of the crystal. The dashed line corresponds to the
expected behaviour of a diffusive motion with a diffusion constant as shown.
\R-vst\
<t>2 = |0 o | 2 sech
Ro
BEC of Excitonic Particles in Semiconductors 353
Comparison of the experimental results with the prediction of the model
reveals excellent agreement [34]. One interesting aspect of the prop-
agation of such a soliton pulse is its velocity, vs = 2.98 x 105 cm/s,
which is significantly smaller than the velocity of sound in the crystal,
v\ = 4.5 x 105 cm/s. Further investigations are needed to understand
more fully the formation and dynamics of the spatially non-uniform
condensate on which our interpretation rests.
5 Conclusions
There is growing evidence that excitonic particles in semiconductors can
exhibit features expected in a weakly interacting Bose gas. Luminescence
and pump/test experiments show features which are explained in a
natural way by the ideal, or nearly ideal, Bose gas model. Anomalous
transport of excitons initially confined close to the crystal surface is
observed. This ballistic transport over remarkably long distances suggests
that a superfluid phase of excitons is involved.
References
[I] Typical values of as and EB range from a few meV in Si or Ge to
a fraction of one eV in alkali or copper halides with corresponding
Bohr radius ranging from several tens of nm to less than one nm.
[2] O. Akimoto and E. Hanamura, Phys. Soc. Japan 33, 1537 (1972).
[3] E. Hanamura, Solid State Comm. 12, 951 (1973); A.L. Ivanov and
H. Haug Phys. Rev. B 48, 1490 (1993).
[4] G.M. Gale and A. Mysyrowicz, Phys. Rev. Lett. 32, 727 (1974).
[5] N. Nagasawa, T. Mita and M. Ueta, J. Phys. Soc. Japan 41, 929
(1976).
[6] D. Hulin, A. Mysyrowicz and A. Antonetti, J. Luminescence 30, 290
(1985).
[7] I.M. Blatt, K.W. Boer and W. Brandt, Phys. Rev. 126 1691 (1962).
[8] S.A. Moskalenko, Soviet Physics Solid State 4, 199, (1962).
[9] L.V. Keldysh and A.N. Kozlov, JETP 27 521, (1968). L.V. Keldysh,
Y.V. Kopaev, Soviet Physics Solid State 6, 2219 (1965).
[10] E. Hanamura and H. Haug, Physics Reports C33, 209 (1979).
[II] C. Comte and P. Nozieres, J. Physique (Paris) 43, 1069 (1982).
[12] L.V. Keldysh in Proc. 9th Int. Conf. on Physics of Semiconductors
(Moscow 1968) p. 1303; see also T.M. Rice, J. Haensel, T. Phillips,
G.A. Thomas, in Solid State Physics 32, 1 (1977).
354 A. Mysyrowicz
[13] F. Bassani and M. Rovere, Solid State Comm. 19, 887 (1976).
[14] L.H. Nosanow, J. de Physique 41 C7-1 (1980).
[15] J.J. Hopfield, Phys. Rev. 112, 1555 (1958).
[16] R.S. Knox, Theory of Excitons (Academic Press, New York, 1963).
[17] For a review see A. Goldmann, Phys. Stat. Sol. B 81 9 (1977).
[18] For a review of the literature on biexcitons in CuCl, see, for instance,
C. Klingshirn and H. Haug, Physics Reports 70 315 (1981).
[19] L.L. Chase, N. Peyghambarian, G. Grynberg and A. Mysyrowicz,
Phys. Rev. Lett. 42, 1231 (1979); N. Peyghambarian, L.L. Chase and
A. Mysyrowicz, Phys. Rev. B 27 2325 (1983).
[20] M. Hasuo, N. Nagasawa, T. Itoh, and A. Mysyrowicz, Phys. Rev.
Lett. 70, 1303 (1993).
[21] For a review of excitonic properties of CU2O see, for instance, S.
Nikitine in Optical Properties of Solids, S. Nudelman and S.S. Mitra,
eds. (Plenum, New York, 1969).
[22] LI. Kreingold and V.L. Makarov, Soviet Physics Solid State 15, 890
(1973).
[23] A. Mysyrowicz, D. Hulin and A. Antonetti, Phys. Rev. Lett. 43,
1123, 1275(E) (1979).
[24] P.D. Bloch and C. Schwab, Phys. Rev. Lett. 41, 514 (1978).
[25] D.W. Snoke, A.J. Shields and M. Cardona, Phys. Rev. B 45, 11693
(1992).
[26] P. Yu and Y.R. Shen, Phys. Rev. Lett. 32, 939 (1974).
[27] D. Hulin, A. Mysyrowicz and C. Benoit a la Guillaume, Phys. Rev.
Lett. 45, 1970 (1980).
[28] D. Snoke, J.P. Wolfe and A. Mysyrowicz, Phys. Rev. Lett. 59, 827
(1987).
[29] D. Snoke, J.P. Wolfe and A. Mysyrowicz, Phys. Rev. B 41, 11171
(1990). See also article by Wolfe et al. in this volume.
[30] V.A. Gergel, R.F. Kazarinov and R.A. Suris, JETP 27, 159 (1968).
[31] B. Link and G. Baym, Phys. Rev. Lett. 69, 2959 (1992).
[32] E. Tselepis, E. Fortin and A. Mysyrowicz, Phys. Rev. Lett. 59, 2107
(1987).
[33] E. Fortin, S. Fafard and A. Mysyrowicz, Phys. Rev. Lett. 70, 3951
(1993).
[34] A. Mysyrowicz, E. Fortin, E. Benson, S. Fafard and E. Hanamura,
to be published; M. Inoue, E. Hanamura, J. Phys. Soc. Japan 41,
771 (1976).
[35] E. Hanamura, Solid State Comm. 91, 889 (1994).
15
Crossover from BCS Theory to
Bose-Einstein Condensation
Mohit Randeria
Argonne National Laboratory
MSD 223, 9700 S. Cass Avenue
Argonne IL 60439
USA
Abstract
This article gives a detailed review of our current theoretical understanding
of the crossover from cooperative Cooper pairing to independent bound-
state formation and Bose-Einstein condensation in Fermi systems with in-
creasing attractive interactions.
1 Introduction
There are two well-known paradigms for understanding the phenomena
of superconductivity and superfluidity:
(I) BCS theory [1], in which the normal state is a degenerate Fermi liquid
that undergoes a pairing instability at a temperature Tc < e^, the
degeneracy scale. The formation of Cooper pairs and their condensa-
tion (macroscopic occupation of a single quantum state) both occur
simultaneously at the transition temperature Tc.
(II) Bose-Einstein condensation (BEC) of bosons at a Tc of the order of
their degeneracy temperature. At a fundamental level these bosons
(for example, 4He or excitons) are invariably composite objects made
up of an even number of fermions. The composite particles form at
some very high temperature scale of the order of their dissociation
temperature TdiSSoc> and these "pre-formed" bosons then condense at
the BEC Tc < Tdissoc.
In most cases of experimental interest the system under consideration
clearly falls into one category or the other. For instance, 3 He is a
Fermi superfluid described by (I) whereas 4He is a Bose superfluid (II).
Essentially all cases of superconductivity in metals that we understand
355
356 M. Randeria
reasonably well are much closer to (I) than to (II), as evidenced by
the existence of fermionic quasiparticles above Tc. Nevertheless, as the
rest of this review will show, it is of great interest to consider models
which interpolate between these two extremes: the weak attraction, high
density limit is described by BCS theory while the strong coupling,
low density limit consists of tightly bound bosons. One outcome of
such a study would be a deeper understanding of the phenomena of
superconductivity, even for systems which are clearly closer to one of the
limits. In addition, there are experimental systems which may be in a
regime intermediate between these two extreme limiting cases, and then
one is forced to study the crossover from collective Cooper pairing to the
formation and condensation of independent bosons.
In the remainder of this section we will first give a very quick review
of the historical development of ideas in this subject, then discuss the
reasons for the recent resurgence of interest in this problem, and conclude
with an outline of the rest of the paper. Readers who would like a preview
of the main results of this article may wish to look at the summary in
the concluding section.
/./ History
The problem of BCS versus Bose-Einstein condensation is an old one.
Some of the earliest attempts in the pre-BCS era to theoretically under-
stand superconductivity in metals were in terms of Bose-Einstein con-
densation, an approach which was perhaps taken farthest by Schafroth,
Blatt and Butler [2], Then came the striking success of the Bardeen-
Cooper-Schreiffer (BCS) theory [1] and superconductivity in metals was
finally understood as a pairing instability of a Fermi liquid state. In
the immediate aftermath of BCS the differences with BEC were empha-
sized : that the Cooper pairs, which were highly overlapping in real space,
should not be thought of as composite bosons, and that the correlations
in the BCS condensate were best described in terms of "momentum space
pairing" rather than "real space pairing" in a highly degenerate Fermi
system.
There are, of course, many similarities in the behavior of Fermi and
Bose superfluids insofar as their macroscopic, coherent properties are
concerned. Off-diagonal long range order (ODLRO) [3] in an appropriate
density matrix is a unifying concept in the study of these two types of
superfluids: the macroscopic occupation of a single quantum state forms
the basis for understanding all bulk superfluid systems. Despite these
Crossover from BCS Theory to BEC 357
similarities between the BCS condensate and the Bose condensate, there
are equally obvious differences at the microscopic level. In particular, the
normal states above Tc are completely different.
Perhaps the first discussion of the possibility of a crossover from
a BCS state to BEC as a function of some parameter, in this case
decreasing carrier density, was by Eagles [4] in the context of a theory
of superconductivity in low carrier concentration systems such as SrTiO3
doped with Zr. This remarkable paper has gone virtually unrecognized
and many results obtained in subsequent papers were, in fact, anticipated
by Eagles.
In a seminal paper, Leggett [5] studied a dilute gas of fermions at
T = 0 with an attractive interaction and showed within a variational
approach that there was a smooth crossover from a BCS ground state
with Cooper pairs overlapping in space to a condensate of tightly bound
diatomic molecules. One of his main motivations was to ask to what
extent one might be able to describe Cooper pairs in superfluid 3He as
giant diatomic molecules. (See also [6-8].)
Some years later this question was taken up by Nozieres and Schmitt-
Rink [9] motivated by the problem of exciton condensation [10] where
it might, in fact, be possible experimentally to go from one limit to
the other by varying the density of carriers. These authors extended
the previous analysis to lattice models and, more importantly, to finite
temperatures. Using a diagrammatic formulation they showed that within
their approximation, the transition temperature Tc evolves smoothly as
a function of the attractive coupling from the BCS to the Bose limit.
The papers of Leggett and of Nozieres and Schmitt-Rink have had
a major influence on all the subsequent work in this field, and we shall
return to a more detailed discussion of these later in this review.
1.3 Outline
The remainder of this article is organized as follows. In Section 2 we
describe two models for studying the BCS-Bose crossover.
In Section 3 we first introduce the functional integral formalism which
will be used in the first part of the paper. Using this we address the
question of pair formation versus condensation. The problems associated
with a finite temperature analysis in two dimensions are mentioned
briefly.
We discuss time-dependent Ginzburg-Landau theory in Section 4, and
describe how the dynamics of the pairs just above Tc evolves from the
BCS to the Bose regime.
We turn to the broken symmetry state in Section 5 focusing mainly
on the T = 0 case. The crossover in the ground state and single-particle
excitation spectrum is discussed using a variational formulation.
In Section 6 we describe the evolution of the collective modes using a
generalized RPA analysis.
In Section 7 we discuss numerical results obtained by the quantum
Monte Carlo method, focusing on recent results on the normal state
correlation functions in the crossover regime in two dimensions.
A summary and conclusions are presented in Section 8.
The main focus of this review will be the work done by the author and
his collaborators (in rough chronological order): J.M. Duan, L.Y. Shieh,
J.R. Engelbrecht, L. Belkhir, N. Trivedi, C. Sa de Melo, R. Scalettar and
A. Moreo. An attempt has been made to reference much of the recent
literature relevant to the models of Section 2 but the author apologizes
in advance to those whose work has been left out.
2 Preliminaries
In this section we introduce two models, one in the continuum and the
other on a lattice, of fermions with attractive two-body interactions. Our
aim is to see how, as a function of increasing attraction, these systems
evolve between two apparently very different regimes (see Fig. 1): BCS
theory in weak coupling with a Fermi liquid normal state and tightly
360 M. Randeria
1
Bose
liquid
Fermi /
liquid^A,
BCS Bose
Coupling
Fig. 1. Schematic phase diagram of the attractive fermion problem. The full
curve is the transition temperature Tc for a continuum model (see text), and the
dashed curve is Tc for the lattice (attractive Hubbard) model. The shaded region
represents a crossover below which pair correlations are important. In the Bose
limit, the crossover scale is the pair dissociation temperature.
bound pairs which Bose condense from a normal Bose liquid in strong
coupling.
The continuum model has the Hamiltonian density
V2 1
where x = (x, T). Decoupling the fermion interaction with the Hubbard-
Stratonovich field A(X,T), which couples to xpxp, and integrating out the
fermions we obtain Z = jDADAexp— Seff[A,A]. The effective action
which defines the 5-wave scattering length as. Recall that as a function
of the bare interaction, l/as increases monotonically from —oo for a very
weak attraction to +oo for strongly attractive interaction. Beyond the
two-body bound state threshold in vacuum (l/as = 0), as is the "size"
of this bound state with binding energy E\> = 1 /ma]. The dimensionless
coupling constant in the dilute gas model is then l/kFaS9 which ranges
from —oo in the weak coupling BCS limit to +oo in the strong coupling
Bose limit.
Crossover from BCS Theory to BEC 363
Using (6) we obtain the equation for the transition temperature in
terms of the renormalized coupling as:
tanh (£k/2T0) 1
m
=£ (7)
n = n o (//,r) = ^ [ l - t a n h l ^ l 1. (8)
k / /
In the weak coupling limit, l/kFas —• —oo and we find the BCS results
fi = eF and To = Se^yn^epQxp (—n/2k F\as\), where y ~ 1.781.
The equations can also be solved analytically in the strong coupling
limit, where we find that the roles of the gap and number equations are
reversed: the gap equation (7) determines \x, while the number equation
(8) determines To. In this limit l/kFas —• +oo and one finds tightly bound
pairs with binding energy Eb = \/ma2s. The non-degenerate Fermi system
has \i ~ -Eb/2 and one finds To ~ Eb/2ln (Eb/eF)y2.
This unbounded growth of the "transition temperature" is an artifact
of the approximation and there is, in fact, no sharp phase transition
at To (outside of weak coupling where To = Tc ignoring the small
effects of thermal fluctuations). The point is that the saddle-point ap-
proximation becomes progressively worse with increasing coupling: the
trivial saddle-point A = 0 can only describe a normal state consisting
of essentially non-interacting fermions. While this is adequate for weak
coupling, in the strong coupling limit unbound fermions are obtained
in the normal state only at very high temperatures. In this limit, where
the system is completely non-degenerate, a simple "chemical equilib-
rium" analysis (boson ^ 2 fermions) yields a dissociation temperature
364 M. Randeria
Tdissoc = Eb/\n {Eb/eF) . The logarithm in the denominator represents
an entropic contribution favoring broken pairs and leads to Tdissoc < Eb,
the binding energy. We thus see that in strong coupling To is related to
the pair dissociation scale rather than the Tc ( < To) at which coherence
is established.
A numerical solution of (7) and (8) yields To smoothly interpolating
between the two limiting cases analyzed above. This temperature repre-
sents a "pairing scale" below which pair correlations become important;
for sufficiently large attraction these correlations describe the formation
of real bound states. For temperatures above the pairing scale To, the
system may be described as essentially consisting of unbound fermions.
Here nk is a Fermi function and zg/ = i2ln/(l are the Matsubara Bose
frequencies. The resulting expression for the thermodynamic potential
Q = Qo - jS"1 ^2qMl lnF(q,i<?/), is identical to that obtained by NSR [9]
by summing particle-particle ladders for the free energy. Following these
authors, it is convenient to rewrite Q in terms of a phase shift defined
by F(q,co ± iO) = |F(q,co) | exp(±i<5 (q,a>)). The new number equation
N = — SQ/d/z is then given by
Fig. 2. The transition temperature Tc (full line) and the crossover scale To (dashed
line) plotted as a function of the dimensionless coupling l/kFas for an attractive
Fermi gas in three-dimensions. The BCS limit corresponds to l/kFas —• — oo and
the Bose limit to l/kFas -> oo. All temperatures are in units of the noninteracting
Fermi energy eF. To is obtained from a solution of the saddle-point equations,
while Tc is obtained from the Gaussian approximation described in the text.
(From Ref. [22].)
p 1 • 1 111 11
1 ' ' ' '' ' ' -
\ :
0
-
V
-1 - \ -
T . . . . 1 . . . . .... i A.-
- 2 - 1 0 1
l/(k r a.)
Fig. 3. The chemical potential fi/eF, obtained from the Gaussian approximation,
plotted as a function of the dimensionless coupling l/kFas for an attractive Fermi
gas in three-dimensions. (From Ref. [22].)
which is nothing but the partition function of a free Bose gas. Thus we
see how the BEC result is obtained in the strong coupling limit. In the
next section we will go beyond the Gaussian approximation and see that
the fourth order term describes a repulsive two-body interaction between
the bosons. As is well-known, this repulsion stablizes the low temperature
phase, but does not substantially affect the condensation temperature.
It may be worth commenting on the fact that the numerical result for
Tc (see Fig. 2) appears to be a non-monotonic function of l/kFas with
a maximum value at intermediate coupling which is slightly larger than
the BEC value. We note that this is where the Gaussian approximation
becomes the most questionable (see next section), and even within this
approximation the numerical accuracy is the least. Nevertheless, this
qualitative feature was also seen in Nozieres and Schmitt-Rink's results
for a somewhat different model with a separable potential. The very
large maximum obtained in some other calculations [27] is certainly an
artifact of the approximate form used for the number equation.
Finally, we must mention two other approaches to the finite temper-
ature crossover problem. Haussmann [28] has generalized the Nozieres-
Schmitt-Rink diagrammatic approach using a self-consistent Green's
function formulation which appears to be promising. Gyorffy and
coworkers [29] have used the coherent potential approximation (CPA) to
study the attractive Hubbard model. This method correctly emphasizes
the phase coherence aspect of the transition. However, while this theory
gives reasonable results in the strong coupling regime, it appears not to
work for sufficiently small attraction, at least in its present form.
The Gaussian piece was already studied above (see (9)), and the coefficient
of the nonlinear term is given below in terms of fermion propagators. It
suffices to set the external frequencies and momenta - the arguments of
b - to zero, so that
Z ^ ^ (16)
Collecting the results of the small |q| and small co expansions, and
transforming to (x, t), we obtain the result
a + b\A (x, t) | 2 - ^ V 2
- idj\ A(x, 0 = 0. (17)
We now discuss the two limiting cases, and the intervening singular point,
in more detail.
In weak coupling, we obtain the well-known results a = N(eF)ln(T/
Tc), b = 7C(3)AT(eF)/87r2T2, c = U(?>)N(eF)eF/\2n2T? and d = (inN(eF)/
8TC) -[1 - i(2Tc/neF)]. By rescaling the order parameter *F = >/2cA, we
obtain the conventional TDGL equation:
(e + | ¥ | 2 - <£LV2 + TGLdt) ¥(x, t) = 0, (18)
where e = \T - Tc\/Tc < 1. The characteristic length scale is £(T) =
<^GL£~ 1/2 with £GL ~ vF/Tc, which is the size of the Cooper pair £pair
(defined in the following section). The width of the Ginzburg region,
defined in the usual way as that range of reduced temperatures e for
which the effects of the fourth order term become important, is very
small, being of order (Tc/eF)4.
Crossover from BCS Theory to BEC 371
The dynamics of *F in the BCS limit is overdamped with characteristic
time scale T(T) = TGL£~\ where TGL = n/STc. This damping arises
from the continuum of fermionic excitations into which a pair can decay.
There is in addition an (9 (Tc/eF) propagating part coming from the real
part of d, since our model does not impose particle-hole symmetry [33].
As the coupling increases, we see from (16) that the coefficient of
the propagating piece grows while that of the damped part diminishes.
The singular point fi(Tc) = 0 separates the regime of damped pair
excitations from that of propagating pairs. For stronger coupling, the
fermionic excitations have a gap in the normal state, and an essentially
propagating mode is obtained at this level of approximation [34]. The
full significance of the singular point at intermediate coupling where
fi(Tc) = 0 is not clear at the present time. Nor is it known whether
the thermally excited modes, ignored in this description, smooth out the
effects of this singularity.
In the extreme strong coupling limit we find the TDGL coefficients to
be given by a = KU (\fi\^2 - (Eb/2)V2), b = KU/32\II\3/2, C = KTT/16|/I|1/2,
and d = Kn/lfi^2, where K = N(eF)/elJ2. Note that we can set \i ~ —E b/2
in b,c and d. Using the normalization ^o = s/dA we can rewrite (17) as
- jW0 + U\¥o\ 2Vo ~ (2M)"1V2^o - id Wo = 0. (19)
This is simply the Gross-Pitaevskii equation for a dilute gas of bosons
of mass M = 2m with a repulsive interaction U = Ana^/M character-
ized by a (boson) scattering length a\> = 2as > 0 with n ^ <C 1, where
nb = n/2 ~ kp (in terms of the kF of the constituent fermions). This re-
pulsive interaction between the composite bosons has also been obtained
independently by Haussmann using a self-consistent Green's function
method [28].
The chemical potential of the bosons controls the phase transition via
the change in sign of ]i = Eb — 2|/x|. The prefactors of the divergent length
and time scales at this transition are given by £GL — kjl/^fkFas > ^pair —
as and TQL — T~l/(kFas) respectively. These are both much longer than
the microscopic scales because of the diluteness condition kFas <C 1.
Consequently the Ginzburg region (s <C kFas) is again small in strong
coupling.
We will find in the next section that the size of the pairs ^pair defined in
terms of the T = 0 wavefunction is a monotonically decreasing function
of the attractive interaction, as one might guess on physical grounds. The
"Ginzburg-Landau" coherence length in the dilute gas model is, however,
a non-monotonic function of the coupling. A numerical evaluation of the
372 M. Randeria
TDGL coefficients shows that £GL decreases with increasing interactions
at first, reaches a minimum value of order of the interparticle spacing fe^ 1
at intermediate coupling, and then increases again attaining the strong
coupling form discussed above.
Closely related to the large Ginzburg-Landau lengths in the two
extreme limits is the small Ginzburg region found in both the BCS
and the Bose limits of the dilute gas model. Thus the neglect of the
fourth order contributions is justified in both limits. However, in the
intermediate coupling regime, where /CJ^GL ~ 1, one does not have a
small parameter which controls the calculation: the Ginzburg region is
of order e = (T—T c)/Tc ~ 0(1). The only reason to trust the intermediate
coupling results, at least qualitatively, is that they smoothly interpolate
between two non-trivial limiting cases. At T=0 the variational principle
provides further justification, as will be seen in the following section.
= const.
k
M2N(6F) (24)
where the "internal pair wave function" is [36] xp^ = Ak/2£k, which is
related to, but distinct from, the wave function cp^ in (22). Evaluating
the matrix element (using r 2 —• V£) we find that in the BCS limit
£pair ~ ^ F / A > kp1, which is the interparticle spacing. With increasing
coupling the pair size decreases monotonically, with a limiting form
£Pair ~ as < kjl in the Bose regime.
There is an amusing feature in the evolution of the gap to single-
particle (fermionic) excitations [4, 5]. This is given by the minimum of
the Bogoliubov quasiparticle energy E^ within the band, namely,
£ k _ / ^ + AY' . (26)
F / A for/i>0
£ g a p
~ \ (/i2 + A^)1/2 for^<0 (2?)
(for a simple way to see this, see Fig. 6 of Ref. [13]). As pointed out
by Leggett [5], this has an important consequence for the case of non-
5-wave pairs. Even if one has nodes in the gap function in the BCS
limit, there is a non-vanishing gap in the composite boson limit which
simply corresponds to the ionization energy of the diatomic molecule.
The reader can find a detailed discussion of the crossover in non-s-wave
superconductors in Ref. [7] (for the 3D case) and Ref. [13] (in two
dimensions).
Crossover from BCS Theory to EEC 375
5.5 Two Dimensions (2D)
It was shown by Miyake [8] and by Duan, Shieh and the author
(RDS) [13] that, in a 2D dilute gas model, the existence of a two-
body bound state in vacuum is a necessary (and sufficient) condition
for a Cooper instability. For purely attractive two-body potentials V(r)
this is not surprising since an infinitesimal attraction leads to a bound
pair in vacuum. However, as emphasized by RDS, when J d2rV(r) > 0
(e.g., hard-core plus attraction, where this integral is infinite), one has
to cross a finite threshold in the attraction before a bound state forms
in vacuum (absence of a medium), and the non-trivial result is that one
has to cross the same threshold for a pairing instability in a degenerate
Fermi system. In addition RDS showed that this result was special to
the s-wave channel: an /-wave bound state in vacuum is not a necessary
condition for a pairing instability with / > 0 in 2D.
The above result is intimately related to the jump discontinuity in the
2D density of states (DOS) at the bottom of the band, or the related
logarithmic singularity in the 2D propagator and T -matrix. The low
energy s-wave T-matrix for two-particle scattering in vacuum (i.e., in the
absence of a medium), which we shall need below, has the form
To(co) = (4n/m) [- ln(co/Ea) + in] ~*, (28)
where Ea is the binding energy of the two-body bound state. The
continuum model in 2D has another feature - a constant DOS - which
greatly simplifies the crossover calculation [4, 8, 13]. For the dilute
gas model in 2D this allows an exact solution of the T = 0 gap and
number equations for all couplings leading to very simple and transparent
results [13].
The variational gap equation 1/g = — ^ k l/2£k, with Ak = A, has an
ultraviolet divergence which is regulated by using the T-matrix. Unlike
the 3D case (6), however, we cannot take the zero energy limit at the
outset since the two-dimensional T-matrix (28) vanishes logarithmically.
We find the "renormalized" gap equation
i
4TT _ r d r Ll
mT0(2co) " Jo £ k L £ k - c o - i O + ~ £ i ; j ' ( )
which is independent of co. The only way in which the attractive interac-
tion enters is via the energy of the bound state in vacuum, which must
exist as described above.
The dimensionless coupling constant in the 2D case - the analog of
1/kpds in 3D - is given by Ea/eF9 which varies from 0 (in the BCS limit)
376 M. Randeria
to oo (in the Bose limit). It is then easy to solve the above gap equation
together with (23) for arbitrary Ea/eF to obtain
This simple analytical result clearly shows a smooth crossover from the
BCS to the Bose limit. Note that the essential singularity of the BCS
limit is buried inside the dependence of the binding energy Ea on the
attractive interaction. For a detailed discussion of the 2D crossover
problem, including an extension to higher angular momentum pairing,
the reader is referred to Ref. [13].
In the T = 0 limit, the above equations clearly reduce to the gap and
number equations derived from the variational approach above.
It is important to emphasize that the saddle-point approximation is
much more reliable for T <C Tc than it was for T > Tc. Outside of the
weak coupling regime, the important physics of pairing correlations, and
eventually bound state formation, above Tc is in fluctuations beyond the
trivial saddle-point. In the broken symmetry state, however, the non-
trivial saddle-point of Seff includes the non-perturbative effects of bound
state formation and condensation. At least for T <C Tc, the fluctuations
about the non-trivial saddle-point make only a small correction even for
strong coupling.
Crossover from BCS Theory to BEC 377
6 Collective Excitations
Gaussian fluctuations about the broken symmetry saddle-point represent
the collective excitations of the superconductor. The functional integral
approach is a very elegant way to study the crossover problem for
collective modes, see Ref. [23]. We present here a simpler derivation
using the old-fashioned RPA [37] based on the work of Belkhir and
the author [35]. We should also mention that related results on the
evolution of the collective modes have been obtained by many authors
for the BCS-Bose crossover [38, 39], for the crossover in the exciton
condensation problem [40], and for the crossover from spin-density wave
(SDW) to local moment antiferromagnetism [18, 41].
The RPA analysis of collective modes has also been extended to
charged systems. While long range Coulomb interactions have been
largely ignored up to this point in our discussion, we discuss them here
since they have a profound impact on the collective modes [37, 42]. At
the end of the section we will briefly mention some results on charged
systems based on Ref. [35]. For a comprehensive treatment of plasmons
in layered systems, especially in the context of the BCS-Bose crossover,
the reader is referred to the recent work of Cote and Griffin [39].
Our study of collective modes is motivated by the following reasons.
The important excitations in the BCS limit involve broken pairs. With
increasing attraction these are pushed to very high energies, and in
the Bose limit it is the collective modes which are the dominant low-
energy excitations. Thus one might expect that a proper description of
the intermediate coupling regime must include both broken pairs and
collective modes.
Our starting point is the attractive Hubbard Hamiltonian (2) written
in momentum space
H
= Y,(£k " ric&ck* ~ M 5 1 ck+qi4-qick'lCkh (33)
M
k,a kk'q
where Sk = —2t J^f=1 cos(fcja) and M is the total number of lattice sites.
The "reduced" Hamiltonian ifbcs includes only the fc' = — k part of the
interaction term in (33), and thus only describes the interaction between
pairs with zero center-of-mass momentum.
The variational solution described above can be recast in the form
of a linearization of the equations of motion for CkG and c£a with the
dynamics governed by ifbcs- This is the well-known Hartree-Fock-
Bogoliubov scheme and one arrives at the familiar number and gap
378 M. Randeria
equations. A and the chemical potential ju are given by the solution
of 1/(7 = (2M)- 1 £ k £ ^ and / = (2M)" 1 £ k (1 - £k/Ek). Here £k =
ek-Ji, Ek = (<^+A 2 ) 1/2 , and / = N/2M is the filling factor. The only real
difference with the continuum model treated in earlier sections is that
one must take into account the Hartree shift in the chemical potential:
/} = // + fU, which can be significant for large U.
To determine the collective mode spectrum we study the time evolution
of density fluctuations c^+q(TCka- Its equation of motion is coupled to
that of pairs with a finite center of mass momentum cjj"+q+c+k\ and
c-k-q[Ck\^ resulting from the particle-hole mixing due to the condensate:
(cfctcifci) ^ 0- The overall strategy of the RPA is simple: first, to linearize
the equations of motion [37] with respect to the mean field ground state,
and, second, to diagonalize the resulting Anderson-Rickayzen equations
by finding the appropriate eigenoperators for the collective coordinates.
At the mean field level one has diagonalized only the H\>cs part of the
full Hamiltonian H = H\>cs + Hmi. At the RPA level, we treat the small
fluctuations introduced by H{nt which describes the interaction between
the Bogoliubov quasiparticles.
The actual implementation of this idea is algebraically messy. We
omit all details, referring the reader to Ref. [35], which generalizes the
weak coupling calculation of Bardasis and Schrieffer [43] to the crossover
problem. The BCS limit is considerably simplified by an approximate
particle-hole symmetry arising from the fact that only fermions within a
thin shell symmetric about the Fermi energy are affected by the pairing.
For arbitrary U, one does not have such a p-h symmetry and the
calculations are more complicated.
To determine the collective excitation spectrum co(q) we take the matrix
elements of the equations of motion between the ground state (which is
the BCS state renormalized by the zero-point motion of the collective
modes), and a state containing exactly one quantum of excitation. The
resulting secular equation is
Here a,fc,c denote any one of the following quantities: the excitation
Crossover from BCS Theory to BEC 379
energy co(q)9 or the quasiparticle energy EkA = Ek+q+Ek, or the coherence
factors l{k,q) = ukuk+q + vkvk+q, m(k9q) = ukvk+q + vkuk+q, and n(k,q) =
ukuk+q — vkvk+q. We now solve (34) for co{q\ restricting attention to the
long wavelength q —> 0 regime.
In weak coupling, the integrals Iw,n,i and IE,n,m vanish as a result of the
p-h symmetry discussed above, and we are left with a 2 x 2 determinant
to solve. The small q and co expansion of the remaining terms in
(34) is conveniently written in terms of four quantities: x = X^^/f 3 >
y = ZkWktf/El w = Zk^U/El and z = A2^k(Wki)2/E5k. The
momentum sums can be readily evaluated to obtain the q —• 0 dispersion
relation in d dimensions
Fig. 4. The normal state spin susceptibility for the 2D attractive Hubbard model
at quarter-filling (n) = 0.5 plotted as a function of temperature, for various U
values. The full curve is for the non-interacting system. The U/t = 4 and 8 data
are from Monte Carlo simulations on an 8 x 8 lattice, and the U/t = 12 data are
on a 4 x 4 lattice. The error bars are of the size of the symbols. The RPA curve
is fit to the U/t = 4 data (not shown in the diagram) in the high temperature
regime 2t < T <4t using a "renormalized" U — 3.25. (From Ref. [14].)
the normal state then clearly arise from the formation of singlet pairs
without any coherence, and not from superconducting fluctuations.
A static susceptibility with dxs/dT > 0, though suggestive, is by itself
not sufficient to establish a "spin-gap", namely a reduction in the low fre-
quency spectral weight Im%(q, co), which is probed by the NMR relaxation
rate \/T\T = lim^oYlq Im z(^co)/co. In general, it is very difficult to
obtain Im/(q, co) from imaginary-time Monte Carlo data, since it requires
analytic continuation from Matsubara to real frequencies. However, we
use a trick that allows us to extract \/T\T directly from the imaginary
time spin-spin correlation function measured at T = 1/2T, provided the
temperature is much smaller than the characteristic frequency of the
spectral weight Im#(q, co) (see Ref. [14] for details).
The relaxation rate \/T\T determined from this analysis is plotted
in Fig. 5, it is found to decrease as T is lowered. In Fig. 6 we show
Crossover from BCS Theory to BEC 385
0.1 1
1 1 1 ' 1
A
(n>=0.5-
A
A U=4 -
A D U=8
A
-0.05 -
A
A •
D
i rh Di 1 1 , 1
0
0 0.25 0.5 0.75
T/t
Fig. 5. The normal state NMR relaxation rate \/T\T (in arbitrary units) for the
2D attractive Hubbard model, at quarter-filling (n) = 0.5, for U/t = 4 and 8 on
8 x 8 lattices. (From Ref. [14].)
that \/T\T indeed tracks the static susceptibility &, thus establishing
spin-gap behavior in the normal state. It is worth noting that, for the
two values of U studied, the data for l/T\T vs. Xs appear to lie on the
same curve.
Is the opening up of a "spin gap" accompanied by a charge gap, or
a pseudo-gap? In this model the answer appears to be yes, although
more detailed work is necessary. The single-particle density of states
obtained via analytic continuation in Ref. [57] gives some evidence for
a gap-like structure developing above Tc. We have also studied the
momentum distribution function (rik). While (rik) is clearly broader than
what would be expected for a Fermi gas (Z = 1) with thermal smearing,
it is difficult to conclude (from fitting the Monte carlo data to various
functional forms) whether the observed behavior necessarily implies a
gap. However the measured (rik) again clearly demonstrates that the
system is in a degenerate Fermi regime.
Finally, we discuss the applicability of our results to normal state
386 M. Randeria
0.1 1
I 1
_ A
A
A
A
-0.05-
A
A (n)=0.5
A U=4
R • U=8 _
D
0 n i 1 , 1
0 0.05 0.1 0.15
Fig. 6. A parametric plot of the normal state NMR relaxation rate 1/TiT (from
Fig. 5) versus the spin susceptibility (from Fig. 4) for the 2D attractive Hubbard
model. For a Fermi liquid all of the data, for a given [/, should collapse to a
single point; instead the two quantities track each other: 1/TiT ~ Xs- (From
Ref [14].)
NMR experiments [49, 50, 51] on the YBa2Cu3O6+x systems; while these
are the best studied systems, spin-gap behavior is not [52] restricted
to them. Two issues (at least) need to be discussed: (1) the role of
antiferromagnetism (AFM), and (2) why spin gaps are seen in some
materials, and not others.
In the Tc = 60K system (x ~ 0.65) the Knight shift, which probes #s,
increases by a factor of four as T increases from Tc to 300K, which we
would identify as Tp. For the O and Y sites (where the form-factors [56]
filter out the AFM contribution) 1/TiT indeed tracks Xs below 300K.
Our model, which has no AFM, is able to explain these spin gap features
naturally. For the Cu site, l/T\ is larger in magnitude and shows spin gap
behavior only below 150K. This would appear to require a combination
of AFM fluctuations and a spin gap opening up due to enhanced pairing
correlations. Clearly our simple model does not have any non-trivial
q-dependence of the spin gap.
Crossover from BCS Theory to EEC 387
In the Tc = 90K (x ~ 1.0) material, spin gap behavior is not seen (or
is seen over a very narrow regime). To discuss the range Tc < T < Tp
over which these anomalies will be seen, one needs to know how Tc
and Tp change as a function of doping (x). Since the attractive Hubbard
model is not a microscopic model for YBCO (e.g., half-filling in this model
has nothing to do with the magnetic insulator at x = 0), we can only
make qualitative remarks. Let us assume that once YBCO is metallic,
increasing x is analogous to increasing the carrier concentration n in the
model. For fixed (7, Tc increases with n, till n ~ 0.85, and then drops to
zero at n = 1.0 in 2D, as shown in Ref. [47]. On the other hand, Tp is
expected to be relatively independent of n for smallrc,to the extent that
one has independent pair formation, and could even drop with increasing
n due to the Pauli principle blocking the formation of independent pairs.
This crude argument suggests why the system with the highest Tc, and
highest carrier concentration, might have the smallest window between
the pairing temperature and Tc.
References
[I] J. R. Schreiffer Theory of Superconductivity (Benjamin/Cummings,
Reading, 1964).
[2] For a review, see J. M. Blatt Theory of Superconductivity (Academic,
New York, 1964) and references therein.
[3] C. N. Yang, Rev. Mod. Phys. 34, 694 (1962).
[4] D. M. Eagles, Phys. Rev. 186, 456 (1969).
[5] A. J. Leggett in Modern Trends in the Theory of Condensed Matter,
A. Pekalski and R. Przystawa, eds. (Springer-Verlag, Berlin, 1980);
and J. Phys. (Paris) Colloq. 41, C7-19 (1980).
[6] N. D. Mermin and P. Muzikar, Phys. Rev. 21, 980 (1980).
[7] M. G. McClure, D. Phil, thesis, University of Sussex (1981).
[8] K. Miyake, Prog. Theor. Phys. 69, 1794 (1983).
[9] P. Nozieres and S. Schmitt-Rink, J. Low Temp. Phys. 59, 195 (1985).
[10] See C. Comte and P. Nozieres, J. Phys. (Paris) 43, 1069, 1083 (1982);
H. Haug and S. Schmitt-Rink, Prog. Quant. Elect. 9, 3 (1984); and
references therein.
[II] J. Bednorz and K.A. Muller, Z. Phys. B 64, 189 (1986); M.K. Wu et
ah, Phys. Rev. Lett. 58, 908 (1987).
[12] See High Temperature Superconductivity, Proceedings of the Los
Alamos Conference, K. Bedell et al, eds. (Addison-Wesley, Redwood
City, Calif., 1990).
[13] M. Randeria, J. Duan, and L. Shieh, Phys. Rev. Lett. 62, 981 (1989);
Phys. Rev. 41, 327 (1990).
[14] M. Randeria, N. Trivedi, A. Moreo, and R. T. Scalettar, Phys. Rev.
Lett. 69, 2001 (1992) and in preparation.
[15] A. S. Alexandrov and J. Ranninger, Phys. Rev. 23, 1796 (1981); S.
Robaszkiewicz, R. Micnas, and K. A. Chao, Phys. Rev. 23, 1447
(1981).
[16] R. Micnas, J. Ranninger, and S. Robaszkiewicz, Rev. Mod. Phys.
62, 113 (1990) and references therein.
[17] See the reviews by H. Capellmann, and by T. Moriya in Metallic
Magnetism, H. Capellmann, ed. (Springer, Berlin, 1986); G. G.
Lonzarich and L. Taillefer, J. Phys. C 43, 4339, 1083 (1985).
[18] J. R. Schrieffer, X. G. Wen, and S. C. Zhang, Phys. Rev. B 39, 11663
(1988).
390 M. Randeria
[19] M. P. A. Fisher, G. Grinstein, and S. M. Girvin Phys. Rev. Lett.
64, 587 (1990); A. F. Hebard and M. Paalanen, Phys. Rev. Lett. 69,
1604 (1992).
[20] K. Moulopoulos and N. W. Ashcroft, Phys. Rev. Lett. 66, 2915
(1991).
[21] M. Schmidt, G. Ropke, and H. Schulz, Ann. Phys. 202, 57 (1990);
T. Aim, B. L. Friman, G. Ropke, and H. Schulz, Nucl. Phys. A 551,
45 (1993); see also G. Ropke, this volume.
[22] C. A. R. Sa de Melo, M. Randeria, and J. R. Engelbrecht, Phys.
Rev. Lett. 71, 3202 (1993).
[23] J. R. Engelbrecht, C. A. R. Sa de Melo, and M. Randeria (preprint,
1993).
[24] V. N. Popov, Functional Integrals and Collective Excitations (Cam-
bridge University Press, 1987); for a nice introduction and references,
see V. Ambegaokar in Percolation, Localization and Superconductiv-
ity, A. M. Goldman and S. A. Wolf, eds. (Plenum, New York,
1984).
[25] For a more careful derivation, see M. Randeria, J. Duan and L.
Shieh, Phys. Rev. 41, 327 (1990).
[26] See, e.g., E. M. Lifshitz and L. P. Pitaevskii, Statistical Physics, Part
I (Pergamon, Oxford, 1980), Section 77.
[27] M. Drechsler and W. Zwerger, Ann. Physik 1, 15 (1992); W. Zwerger,
(preprint 1993).
[28] R. Haussmann, Z. Phys. B 91, 291 (1993), and private communica-
tion.
[29] B. L. Gyorffy, J. B. Staunton, and G. M. Stocks Phys. Rev. B 44,
5190 (1991).
[30] S. Schmitt-Rink, C M . Varma and A. Ruckenstein, Phys. Rev. Lett.
63, 445 (1989); J. Serene, Phys. Rev. B 40, 10873 (1989); A. Toku-
mitu, K. Miyake, and K. Yamada, Prog. Theor. Phys. Suppl. 106,
63 (1992); Phys. Rev. B 47, 11988 (1993).
[31] P.J. H. Denteneer, G. An, and J. M. J. van Leeuwen, Europhys. Lett.
16, 5 (1991).
[32] E. Abrahams and T. Tsuneto, Phys. Rev. 152, 416 (1966); see also
M. Cyrot, Rep. Prog. Phys. 36, 103 (1973) and references therein.
[33] H. Ebisawa and H. Fukuyama, Prog. Theor. Phys. 46, 1042 (1971).
[34] Even in a purely bosonic system the dynamics of the order parameter
will be damped sufficiently close to Tc (see P. Hohenberg and B.
I. Halperin, Rev. Mod. Phys. 49, 435 (1977), section VI E). Our
Crossover from BCS Theory to BEC 391
analysis, however, is restricted to be valid only outside the critical
region.
[35] L. Belkhir and M. Randeria, Phys. Rev. B 45, 5087 (1992); Phys.
Rev. B 49, 6829 (1994).
[36] See, e.g., A. J. Leggett in Quantum Liquids, J. Ruvalds and T. Regge,
eds. (North Holland, Amsterdam, 1978), p. 167.
[37] P. W. Anderson, Phys. Rev. 112, 1900 (1958).
[38] J. Sofo, C. A. Balseiro, and H. E. Castillo, Phys. Rev. B 45, 9860
(1992); T. Kostryko and R. Micnas, Phys. Rev. B 46, 11025 (1993);
A. S. Alexandrov and S. G. Rubin, Phys. Rev. B 47, 5141 (1993).
[39] R. Cote and A. Griffin, Phys. Rev. B 48, 10404 (1993).
[40] S. Schmitt-Rink, D. S. Chemla, and H. Haug, Phys. Rev. B 37, 941,
(1988); R. Cote and A. Griffin, Phys. Rev. B 37, 4539 (1988).
[41] H. Monien and K. Bedell, Phys. Rev. B 45, 3164 (1992); P.J. H.
Denteneer and J. M. J. van Leeuwen, Europhys. Lett. 22, 413 (1993).
[42] For a review and references, see P. C. Martin in Superconductivity,
R. D. Parks, ed. (M. Dekker, New York, 1969), Part 1, p. 371.
[43] A. Bardasis and J. R. Schrieffer, Phys. Rev. 121, 1050 (1961).
[44] L. Foldy, Phys. Rev. 125, 2208 (1962).
[45] See the review by D. J. Scalapino in High Temperature Supercon-
ductivity, Proceedings of the Los Alamos Conference, K. Bedell et al,
eds. (Addison-Wesley, Redwood City, Calif., 1990).
[46] R. T. Scalettar et al, Phys. Rev. Lett. 62, 1407 (1989).
[47] A. Moreo and D. J. Scalapino, Phys. Rev. Lett. 66, 946 (1991).
[48] J. E. Hirsch, Phys. Rev. B 28, 4059 (1983).
[49] W. W. Warren et al, Phys. Rev. Lett. 62, 1193 (1989); R. E. Walstedt
et al, Phys. Rev. B 41, 9574 (1990).
[50] M. Takigawa et al, Phys. Rev. B 43, 247 (1991).
[51] H. Alloul, T. Ohno, and P. Mendels, Phys. Rev. Lett. 63, 1700 (1989).
[52] H. Zimmerman et al, Physica C 159, 681 (1989); R. E. Walstedt, R.
F. Bell and D. B. Mitzi, Phys. Rev. B 44, 7760 (1991).
[53] J. Orenstein et al, Phys. Rev. B 42, 6342 (1990); Z. Schlesinger et
al, Phys. Rev. Lett. 67, 2741 (1991); C. C. Homes et al, Phys. Rev.
Lett. 71, 1645 (1993).
[54] R. Liu et al, Phys. Rev. B 46, 11056 (1992).
[55] In 2D where the Kosterlitz-Thouless Tc is much lower than Tc°, it
is interesting to ask if deviations from Fermi liquid behavior persist
even as U —> 0. Finite size errors, which are more pronounced at
small U, make it difficult to address this question via Monte Carlo
calculations.
392 M. Randeria
[56] B. S. Shastry, Phys. Rev. Lett. 63, 1288 (1989); A. J. Millis, H.
Monien, and D. Pines, Phys. Rev. B 42, 167 (1990); N. Bulut. D.
Hone, D. J. Scalapino and N. E. Bickers, Phys. Rev. B 41, 1797
(1990).
[57] A. Moreo, D. J. Scalapino, and S. R. White, Phys. Rev. B 45, 7544
(1992); A. Moreo, N. Trivedi, and M. Randeria (unpublished).
16
Bose-Einstein Condensation of
Bipolarons in High-Tc Superconductors
J. Ranninger
Centre de Recherches sur les Tres Basses Temperatures,
CNRS, BP 166
38042 Grenoble-Cedex 9
France
Abstract
The short coherence lengths and the low carrier concentrations in high
temperature superconductors (HTCSC) quite naturally favor a scenario
where preformed bound electron pairs undergo a Bose-Einstein conden-
sation (BEC). There are numerous experimental indications in the normal
state properties (anomalous diamagnetic susceptibility, NMR spin-lattice
relaxation, Knight shift and crystal electric field transitions, as well as
entropy non-linear in T and strong local lattice deformations seen by EX-
AFS), all of which indicate a characteristic temperature TPB associated
with the breaking of such preformed electron pairs and the existence of
a pseudogap above Tc. As the number of charge carriers is increased by
chemical doping, TPB decreases and the critical temperature Tc, where su-
perconductivity sets in, in this so-called "underdoped regime", increases.
The maximum value for Tc is reached when Tc < Tp#. Upon further dop-
ing, going into the so-called "overdoped regime", Tc decreases. We take this
as an indication that the superconducting state depends on the existence of
preformed electron pairs. The normal state properties of the "underdoped
regime" seem to show significant differences from Fermi liquid behavior,
while in the "overdoped regime" they seem to be typical of ordinary metals.
In order to capture this empirically established scenario in HTCSC, we
discuss a phenomenological model based on a mixture of localized bosons
(bound electron pairs such as bipolarons) and itinerant fermions (elec-
trons), with a local exchange between bosons and fermion pairs. Super-
conductivity in such a system materializes when, due to the boson-fermion
exchange mechanism, a BCS-like gap opens up in the electron spectrum. At
the same time, the initially localized bosons acquire itinerancy and condense
inside the energy gap of the fermions. The dependence of Tc on the total
393
394 J. Ranninger
number of carriers is reminiscent of the BEC temperature of lattice bosons,
with an effective mass determined by the boson-fermion pair exchange term.
Above Tc, the bosons behave as resonant states inside the Fermi sea, which
is manifest by the appearance of a pseudogap and a blocking of the chem-
ical potential close to the bosonic level
1 Introduction
Since the experimental discovery of the first superconducting material
(Hg) by Kamerlingh Onnes in 1911 [1], progress in producing super-
conducting materials with critical temperatures higher than about 20 K
stagnated until the discovery of the high Tc superconductors by Bednorz
and Muller and Chu et al. [2].
As far as the classical (pre-high Tc superconductors) are concerned,
they could be described by the BCS theory [3] based on the idea of
pairing of electrons in k-space - called Cooper pairs. Within the weak
coupling limit of this theory, the critical temperature Tc is given by:
Tc ~ 1.13co0e~1Aeff, (1.1)
where Aeff = \i — K \£ denotes the effective Coulomb repulsion and
X is the phonon-mediated electron-electron attraction. The prefactor
in the expression for Tc being proportional to a>o - the characteristic
phonon frequency of the material - perfectly describes the isotope effect
Tc oc M" 1 / 2 (M being the ionic mass), thus confirming the phonon-
induced origin of electron pairing.
It was quickly understood [4] that, on the basis of the weak coupling
theory of superconductivity, Tc could not be expected to exceed 30-40 K.
The reason for this lies in the interdependence of the parameters entering
this theory. The effective attraction given by Aeff comes about via a
subtle decrease in efficiency of the Coulomb repulsion \i —> /i* due to
retardation effects. In fact, the quantity
evident from the photo-emission data [20] which measure abrupt changes
of the number of Cu + + (1) atoms as a function of chemical doping and
also confirm equally abrupt increases of the overlap of electronic wave
functions between the Cu(2) ions in the CuC>2 planes and the O(4) apex
oxygens [21]. The reason for the ordering of the chemically doped atoms
leading to an homogeneous distribution of bipolaronic sites and the onset
of a metallic state must lie in the energy gained by the bipolarons due
to delocalization in the normal state and the possibility of achieveing a
macroscopic occupation of the q = 0 state in the superconducting state
due to a BEC of bipolarons. Judging from the EXAFS measurements,
this charge transfer is of a dynamical nature [16,17], leading to a situation
where a bipolaron on any arbitrary site of the atomic clusters involving
the apex oxygens can be exchanged by a pair of electrons in the adjacent
CuC>2 planes. This charge transfer can be described by:
(2.1)
where i\^ in (2.1) represent two adjacent molecular units on the apex
oxygen network and 7*1,2 two adjacent Cu(2)-2O(2)-2O(3) clusters in
the O1O2 planes (see Fig. 1, for a likely realization of this situation in
metallic YBa2Cu3O6+x). The bipolarons are expected to exist in the form
BEC of Bipolarons in High-Tc Superconductors 401
of intersite bipolarons bfii2 = c+ cJ consisting of two polaronic charge
carriers:
cl = ct\ct)(x-xo))i, (2.2)
(3.1)
The charge transfer between bipolarons and electron pairs has strength v
and the common chemical potential ensures conservation of charge car-
riers; Q denotes the normalization volume. The energy of the bipolaronic
level is given by A, which should roughly correspond to TPB-
We note that, already, the simplest case of a non-interacting boson-
fermion mixture (v = 0) leads to deviations from the BEC of a pure Bose
gas. To show this, we consider free bosons with a finite dispersion Ek =
^ + A — 2/z and free fermions with a dispersion ek = £% — n (see Fig. 2)
with £f = h2k2/2mF and £jf = H2k2/2mB, where mF(B) denotes the fermion
402 J. Ranninger
Fig. 2. Schematic plot of the fermion (ek) and boson (Ek) dispersion for the
non-interacting boson-fermion model.
mB
mF )fjT—*dee^n (e + A —
B 2\i) (3.2)
n = nc 2g (kBTBEc)2 — • (3.4)
(a)
Fig. 3. (a) TBEC as a function of the total carrier concentration n in the non-
interacting boson-fermion model, (b) The condensed fraction no(T) of bosons as
a function of temperature in the non-interacting boson-fermion model (continu-
ous line) and the pure Bose gas (dashed line).
k-q "7
(3.5)
where gF(co) denotes the density of states of the fermions. The width
r>(k, co) is different from zero for Eo < co < EQ+II. The real part RF(k, CO)
of ZF(k,co) is obtained in the usual way via a Hilbert transformation
of FV(k,co). The renormalized spectrum cok of the fermions is then
determined by cok — e^ — RF(k9co\i) = 0, given by the set of intersections
of co — ek with RF(k, CO). AS long as Eo > 0, there always exists a wave
vector kF such that for k < kF, we have cok < 0 and for k > kF, we
have cok > 0. This determines the renormalized Fermi vector kF. As
Eo approaches zero, kF increases and a situation is finally reached for
which no kF can be found within the Brillouin zone for which cok>kF > 0.
This signals the breakdown of the Fermi liquid behavior, manifest in the
absence of a discontinuity of ^(cok).
The limit Eo —• 0 is subtle in the sense that riB(E q) in the expression
for £jr(k,a>) diverges for q —• 0. Rewriting riB(Eq) as:
. (3.8)
]l + nov2 (3.9)
in the fermion Green's function and a corresponding Fermi distribution
function of the form:
The expressions (3.9) and (3.10) have exactly the BCS form for the
superconducting ground state but with a gap proportional to the number
of condensed bosons no.
Let us now turn to the properties of boson Green's function in the
same limiting case, i.e. T —• 0 and Eo —• 0. From the discussion of the
instability of the fermionic system, we know that it is triggered by the
macroscopic occupation of the q = 0 mode. The imaginary part of the
BEC of Bipolarons in High-Tc Superconductors 405
(a) (b)
Fig. 4. Fermionic (a) and bosonic (b) self-energy diagrams in terms of fully
dressed self-consistent fermion and boson Green's functions.
Eo = * f" degF(e)—
)€ tanh he - |i). (3.13)
2 J I1 A
In the limit T -> 0, the integrand on the rhs of (3.13) diverges for e = \i
unless gF(e = ix) = 0. This implies the opening up of a gap in the one-
particle electron spectrum which is consistent with the conclusions drawn
earlier from the properties of the fermion Green's function, (3.6)-(3.9).
The arguments developed above clearly demonstrate the mutually
induced superconducting state in the boson-fermion mixture being due
to a BEC of the bosons which occurs with the simultaneous opening of
a gap in the excitation spectrum of the fermions. A self-consistent mean
field theory of this problem has previously been given [24,25].
A fully self-consistent treatment of this problem in the present scheme -
taking into account, in addition to the self-energy diagrams for the normal
phase (Fig. 4), all those including anomalous fermion and boson Green's
functions - will be discussed elsewhere [23]. The critical temperature
406 J. Ranninger
o.io
(a)
Fig. 6. Schematic plot of (a) the boson occupation number n?(T) and (b)
the fermion occupation number n£(T), for the fully self-consistently treated
interacting boson-fermion model.
values for \U\/W describe the "underdoped regime" with a normal state
between Tc and 7># which shows many of the anomalous normal state
properties [38], listed above, and which in many ways are reminiscent of
a purely bosonic system [39].
Concerning the physical foundations of the boson-fermion model, we
propose the following decisive experiments which would either verify or
disprove the relevance of this model:
(1) A verification of the disappearance of polaronic effects upon entering
the overdoped regime for T > Tc using resonant Raman spectroscopy,
photo-luminescence, photo-induced Raman measurements, the appear-
ance of polaron-induced local lattice modes and EXAFS, measuring
polaron-tunneling induced local fluctuations of intramolecular dis-
tances.
(2) A verification of the disappearance of the diamagnetic contribution to
the magnetic susceptibility (not due to superconducting fluctuations)
upon entering the "overdoped" regime. Tendencies suggesting this
disappearance have been observed [29].
(3) A detailed analysis of the doping dependence of the chemical potential
fi by photo-emission spectroscopy. According to the boson-fermion
model, \x is expected to be pinned by the presence of the bosonic
resonant state inside the Fermi sea. Our preliminary calculations on
this issue suggest a rather weak doping dependence of //, provided
the pairing energy A does not vary with doping. Experimentally, this
question is unsettled. There are experiments in which \i is practically
unchanged with doping [40] and those in which n behaves more
like doped semiconductors [41], showing rather significant changes
with doping. High resolution photo-emission experiments testing the
dependence of fi on the carrier concentration and on the temperature,
as well as the existence of a pseudogap, would be strong tests for the
existence of bosonic states (pre-existing electron pairs) above Tc.
Let us now turn to the properties of the superconducting state. Exper-
imentally there are a number of features [35] which clearly deviate from
ordinary BCS-like behavior, such as:
(1) Unusually strong effects of the superconducting state on the lattice
properties. Apart from the classical shift of phonon frequencies with
energies above and below the gap energy [42], one observes strongly
correlated motion of the molecular vibrations (seen by ion channeling
experiments [43]) and a freezing of the vibrational energy of individual
BEC of Bipolarons in High-Tc Superconductors 411
atoms (seen by resonant neutron scattering [44]) which are triggered
by the onset of the superconducting state [45].
(2) An unusually large gap (2A/ks Tc ~ 5 — 8) which in practice does not
vary with temperature and whose signal fades away as one approaches
Tc from below. This is seen from tunneling spectroscopy [46], EELS
[47] and reflectivity measurements [48]. These lattice effects may be
of importance in limiting the fluctuations of the purely electronic
condensation.
(3) An essentially frequency-independent optical conductivity (for frequen-
cies less than the superconducting gap) which shows the same temper-
ature behavior as 1/TiT [49].
(4) Specific heat of the purely electronic contribution which scales sur-
prisingly well with that of 4 He over a wide temperature regime
0 < \T - Tc\/Tc < 0.2 [50], as shown in Fig. 7. The scaling of
the specific heat of HT CSC to that of liquid 4 He permits one to deter-
mine the number of "effective bosons" n& in such systems. The number
derived in this way compares well with the number of charge carriers
determined independently by Hall measurements [51]. Using the for-
mula for the Bose condensation temperature for quasi two-dimensional
systems,
u T2D+e _ h2nnd
B BEC l j
mab ln(2kB Tf^^lK)
permits one to derive the effective mass of the "effective bosons" mab
in the CuC>2 planes. Here mc denotes the "effective boson" mass and d
the size of the molar volume along the c direction. Upon inserting into
(4.1) the experimentally determined transition temperature for T^Q
and for nB the "effective boson" density derived from the comparison
of the specific heat of several HT CSC with that of liquid 4 He, one
obtains mab ~ 3-5 free electron masses if we assume mc/mab ~ 100.
This value agrees reasonably well with the effective masses deduced
from optical experiments [54]. Finally, using the expression for the
penetration depth of charged bosons
= ymabc2/16nnBe2 (4.2)
for the magnetic field orthogonal to the CuC>2 planes, and inserting
the values for mab and nB derived above, one obtains values such as
^H,C ~ 2500-5000 A which are quite compatible with those derived by
independent measurements [51,52].
412 J. Ranninger
40
Bi 2 S r 2 C a 2 C u 3 0 x ,T c = 107K
32 Y B a 2 C u 3 0 7 . 5 , T c = 91.9K
Y Ba 2 Cu 3 O 6 9 7 ) Tc = 92.5 K
1 24
Liquid He , T c = 2 . 1 7 K
O 16
5 Conclusions
The origin of high Tc superconductivity is presently still a matter of
great controversy. Certainly, in these materials we are faced with strong
correlations between the electrons, which most theorists consider as
crucial to explain high Tc superconductivity in terms of a non-lattice-
driven mechanism. As we have stated, there is extensive experimental
evidence for strong coupling of electrons to certain local lattice modes
and consequent polaron formation. On the basis of these experiments,
we are led to a picture in which localized bipolarons form one of the
two subsystems of HTCSC, whereas the strongly correlated electrons
are confined to the other subsystem. A charge exchange, involving
bipolarons and electron pairs, is triggered when one exceeds a certain
doping concentration. At that point, an insulator-metal transition sets
in, monitored by the binding of extra dopant atoms in those materi-
als. The metallic regime consists of two apparently physically different,
regions:
nant state of bipolarons inside the Fermi sea. With increased doping, the
binding energy TPB of such a bipolaronic state decreases and eventually
coincides with the superconducting transition temperature Tc. At that
point, the maximum value of Tc is reached. Upon further doping, Tc
starts to decrease. This whole scenario has been described in terms of
a model of a boson-fermion mixture with an exchange of boson <-•
fermion pairs. Preliminary results of our theoretical investigation of this
model have been presented here [23].
Most of our theoretical arguments for developing the boson-fermion
scenario for HTCSC were based on experimental evidence coming from
the hole-doped cuprates. There is recent experimental evidence that
polaronic charge carriers also exist in electron-doped cuprates [58] as
well as in fullerenes [59]. This strengthens our conviction that the whole
high Tc superconductivity phenomenon has a unique mechanism, such
as the one described here.
Of course, for a satisfactory theory of high Tc superconductivity, the
problem of strong electron-phonon coupling has to be solved in con-
junction with the strong electron correlation problem. One of the key
questions to be addressed in this connection would be a careful study
of the correlation between the pseudogap exhibited in the electron spec-
trum and the pseudo spin gap seen in the magnetic susceptibility of
the insulating and superconducting cuprate HTCSC. The latter has been
extensively studied by inelastic neutron scattering [60] and NMR [61].
The microscopic origin of this spin gap and its evolution in the highly
doped superconducting samples is presently not understood. In princi-
ple, any formation of singlet pairs would produce such a pseudo spin
gap, whatever its origin might be: magnetic, purely correlation driven,
negative U-centers or bipolarons, etc. A prime question to answer
would be to check if a pseudogap in the electron spectrum exists in
the non-magnetic HTCSC. If so, then it would be conceivable that it is
the existence of such a pseudogap which is at the origin of the pseudo
spin gap in systems with magnetic correlations. Such a verification
would be decisive for our understanding of the origin of pairing in
HTCSC.
Abstract
1 Introduction
418
The Bosonization Method in Nuclear Physics 419
V(r)
/
, /
s
/
Fig. 1. The self-consistent potential V(r) in nuclei. The single-particle levels are
labelled by the spectroscopic notation / 7 .
HF = E o (1)
where ep is the single particle energy and vpp> the residual interaction
between particles, p and pf. The Hamiltonian (1) describes a "shell-
model". Introducing a label i to characterize the single-particle states,
420 F. Iachello
one can write a second quantized Hamiltonian
S]
(4)
if ' ^ '*
The Bosonization Method in Nuclear Physics 421
l7
/2
(MeV)
-0.5
- 1.0 -
- 1.5 -
m,m'
(5)
m,m'
m,m'
One may note that the angular momentum zero pairs, called 5 pairs, are
similar to the Cooper pairs of superconductivity
k
is, when written for the finite system,
j,m>0
which include (8) but are more general. If S,D and G pairs are retained,
one has
(S) (D) (G) | 0). (12)
The broken pair states are those containing explicitly the fermion op-
erators cftj. Of course, in constructing the space (13) one has to be
careful about overcompleteness and orthogonality (called the decoupling
problem) but this problem is important only for very small systems.
Once fermion pairs have been introduced, the next step is to map fermion
pair states into boson states [4]. Mapping is a one to one correspondence
in which to each fermion pair state there corresponds a boson state.
When S and D pairs are retained, the mapping is done through the
introduction of s and d boson operators, s,d^ (in = 0,±l,±2), satisfying
Bose commutation relations
=0. (14)
The fermion pair states (11) are mapped into boson states
_L (15)
Explicitly, the first few fermion pair states and their mappings are:
424 F. Iachello
n = 0, v = 0 |0) JV = O, nd = 0 | 0)
n = 2, v= 0 S f | 0) N = 1, nd = 0 s1" | 0)
v= 2 D t 10) Md = 1 dt | o)
n = 4, t> = 0 5 t 2 | 0) JV = 2, nd = 0 s*2 | 0)
i? = 2 S ^ | 0) nd = l s V | 0)
0) nd = 2 rft2 | o)
(16)
The mapping is such that the number of bosons N is half the number of
fermions
N -- n (17)
=
2
and the number of d-bosons is half the fermion generalized seniority
V
nd ==
• (18)
2
As mentioned above, some subtle problems arise in the construction of
the pair space, requiring the introduction of projection operators to take
into account the correct orthogonality properties of the states. These
projection operators are denoted by <PV in Eq. (16). The truncation and
subsequent mapping procedure is shown schematically in Fig. 3.
Proton levels
(82 )
E (MeV) ȣf
0 - 3S
./2
2d
3/2
-1.67 - 2d
5/2
-348 -
^7/2
)
<»
Fig. 4. The valence shells of nuclei between nucleon numbers 50 and 82.
(i) Determine the pair structure constants a,-, /J//, which appear in the
correlated pairs, Eq. (4). Several methods have been suggested, the
simplest one being that of diagonalizing HF in the two-particle valence
space. This method gives the structure constants a,, fijf, and the
single-particle boson energies, eT.
(ii) Determine the boson-boson interaction uxx>m'. Again, several methods
have been suggested, the simplest one being that of equating matrix
elements of the fermion Hamiltonian, HF, in the fermion space of
the left-hand side of Eq. (16), with matrix elements of the boson
Hamiltonian, HB, in the boson space of the right-hand side of Eq.
(16), schematically written
(HF) = (HB) (20)
for four particles (two bosons). This method is referred to as OAI map-
ping [4]. The boson-boson interaction appears here as the contribution
of two terms, the direct and the exchange terms. The exchange term,
which is very important for finite systems, arises from the composite
nature of the bosons (Pauli principle).
At the end of this procedure one has replaced a strongly interacting finite
fermion system by a strongly interacting boson system. The bosonization
of a strongly interacting finite fermion system is also called the Interacting
Boson Approximation (IBA) and it leads to a model of nuclei called the
426 F. Iachello
Interacting Boson Model (IBM) [5]. Two remarks are in order here.
First, the method discussed here is a generalization of methods used in
superconductivity. Superconductivity, when written in the language of
Sections 2-5, corresponds to retaining only S pairs, to considering only
J=0 pairing interactions and to neglecting number projection, hereby
assuming that the structure constants of the S pairs are given by
In Eq. (21), 8j are the single fermion energies, X is the Fermi energy and
A the pairing gap.
The second remark is that the bosonization method can be applied
not only to particle-particle fermion operators, where the product eft eft
is replaced by a boson operator b\ but also to particle-hole fermion
operators [6], where the product eft a is replaced by a boson operator, tf.
The equivalent of the bosonization method then becomes, in the infinite
system, the theory of excitons.
o O<G'
Here sa are the single-particle boson energies and the vaa< are the boson-
boson interactions. A second quantized version of (22) is given in (19)
where T labels the single particle boson states. It is convenient here to
use a spectroscopic notation T = s,d,g,.... The Hamiltonian (22) or (19)
is that of a "shell-model" for bosons. As mentioned in Section 2, in
nuclei, in first approximation, only s and d bosons are important. For
this reason, considerable effort has gone into the study and solution of a
system composed of s and d bosons over the last 20 years.
The Bosonization Method in Nuclear Physics All
3.2 The s-d Boson System
The s-d boson system, usually called interacting boson model [5], is
characterized by the Hamiltonian
J L J
Jo
1
(0)
(23)
0
This Hamiltonian is the most general Hamiltonian invariant under rota-
tions up to two-body boson-boson interactions. Although in principle,
the bosonization procedure, starting from a fermion two-body Hamilto-
nian may produce a boson-boson Hamiltonian with three-body, four-
body, ... terms, in practice it has been found that the higher order terms
are small and therefore it is sufficient to consider the Hamiltonian (23).
One is then asked to find the eigenvalues of HB for a system of N bosons,
where N is the number of active pairs. Several methods have been used
to find the spectrum of (23):
E
(MeV)
56 B0 76
3 -
Exp Th
2 - 6 ^
\ ^j
1 -
°
2
-—I
2 2
0 - o o
Fig. 5. Ab initio calculation of the excitation spectrum of 132Ba using the bosoniza-
tion method. On the left the experimental spectrum, on the right the calculated
spectrum.
(24)
generate a Lie algebra (S. If the index a goes from 1 to n, the Lie algebra
is u(n). For the s-d boson system in which the index a goes from 1
to 6, the Lie algebra is ^ = M(6). The six components are the single
component of the s-boson and the five components of the d-boson, d^
(fi = 0,±l,±2). The Hamiltonian HB can be written in terms of the
operators Ga/j, as
(25)
All the powerful techniques of group theory can then be used to find
the eigenvalues of HB. This is the celebrated "Algebraic Theory" which
has been used extensively in recent years to solve many-body problems.
Algebraic theory is presented in some detail in Ref. [5] and will not be
discussed further here.
The Bosonization Method in Nuclear Physics 429
E
(MeV) 48Cd62 E«p. Th.
2- "A-
4
0*-
2*—
e
4
~CL>
0*—
~2*- 0*— -a-_
'"
2*— 2*-
This method is not used much in practice in nuclei. I will present it here
in view of its relation to the subject of Bose condensation discussed in
this volume, and in view of its interest for systems with N large. I will
begin with the trivial case in which there is no boson-boson interaction.
In this case, the ground state is the state in which all bosons occupy the
lowest level (s-level)(called spherical Bose condensate),
-F=(S^)N I 0) • (26)
y/N\
Excitations are obtained by removing one boson from the lowest level
and placing it into the d-level
0), (27)
The original basis can then be transformed into a deformed boson basis,
obtained by acting with the operators of (30) on a vacuum state
^)bl(p,y)---\O). (31)
The number operator, when written in terms of the new operators
becomes
N = b\bc + b\bfi + b]by + b\bx + b\by + b%. (32)
The ground state is the state of Eq. (29),
N
\0)=\N;P,y). (33)
The Bosonization Method in Nuclear Physics 431
E
(MeV)
8—
6*—
P
Fig. 7. An example of a deformed nucleus: 156Gd.
(34)
where now the indices n and v denote protons and neutrons. One can
consider in principle many pair states, but in practice, in nuclei, only s
and d pairs are important.
4.3 Bosonization
The pair space is then mapped onto a boson space composed of proton
bosons and neutron bosons s\.,d\^ ; sJ,dJjAi . One then obtains the boson
The Bosonization Method in Nuclear Physics 433
Fig. 8. The valence shells for protons and neutrons between particle numbers
50 and 82 (a), and the boson model that replaces the fermion problem after
truncation and mapping (b).
Hv (39)
Methods of Bose condensation can also be used here. The ground state
of the (deformed) two-fluid system is written as
| g) = (Nn \NV l)-1/2(btjN-(bljN* | 0) (42)
and excitations are constructed.
" 1 1' 1 1
s d d
r
> • •
> -
In this case one takes advantage of the fact that the bilinear products of
creation and annihilation operators for bosons and fermions generate a
graded Lie algebra composed of the set of operators of the type
G(B) = })\b
If n is the dimension of the bosonic space (6 for the s-d system) and m is
the dimension of the fermionic space, the (m + n)2 operators of Eq. (44)
generate the graded algebra u(n/m). The powerful machinery of group
theory is then used to help find solutions of the eigenvalue problem for
a mixed system of boson and fermions.
The treatment of mixed systems of bosons and fermions using the co-
herent state method is more complex than the corresponding case of
bosons. The reason is that, while it is possible to introduce in the bosonic
436 F. Iachello
coherent state (29) variational parameters which are c-numbers, when
one deals with fermions such an introduction is not possible. The varia-
tional parameters become Grassman variables in the fermionic case. As
a result, the use of coherent state methods for fermions has not been
developed much.
6 Conclusions
In this article I have presented a brief survey of the bosonization method
in finite systems. The literature on this subject is very extensive and some
references have been indicated in Sections 2-5. Others can be found in
the reference list of Refs. [5] and [11] and in Ref. [12], from which most
of the material reviewed here has been taken. The applications of the
method have up to now been confined to the structure of atomic nuclei
where the method has prove to be very useful. The method can be used
for other finite Fermi systems. An area where, in principle, it could be
very useful is that of the newly discovered metallic clusters. These, too,
are finite Fermi systems, composed of valence electrons. The number of
valence electrons is of the same order of magnitude as the number of
nucleons in nuclei, ~ 100. The residual interaction in metallic clusters
may be different from that of nuclei and therefore the corresponding
boson system may not be an s-d system. None the less, the bosonization
method is sufficiently general that it can be applied to these situations as
well. Work in this direction is in progress [13].
References
[1] C. Racah, Phys. Rev. 63, 367 (1943).
[2] Y.K. Gambhir, A. Rimini and T. Weber, Phys. Rev. 188, 1513
(1969); B. Lorazo, Nucl. Phys. A 153, 255 (1970); I. Talmi, Nucl.
Phys. A 172, 1 (1971). See also I. Talmi, Simple Models of Complex
Nuclei, (Harwood, Chur, 1993) and references therein.
[3] J. Bardeen, L.N. Cooper and J.R. Schrieffer, Phys. Rev. 108, 1175
(1957).
[4] T. Otsuka, A. Arima and F. Iachello, Nucl. Phys. A 309, 1 (1978).
[5] For a review see F. Iachello and A. Arima, The Interacting Boson
Model, (Cambridge University Press, 1987).
The Bosonization Method in Nuclear Physics 437
[6] H. Feshbach and F. Iachello, Phys. Lett. 45B, 7 (1980).
[7] A. Bohr and B.R. Mottelson, Physica Scripta 22, 468 (1980); A.E.L.
Dieperink, O. Scholten and F. Iachello, Phys. Rev. Lett. 44, 1747
(1980); J. Ginocchio and M.W. Kirson, Phys. Rev. Lett. 44, 1744
(1980).
[8] A. Leviatan, Phys. Lett. 143B, 25 (1984).
[9] N.N. Bogoliubov, J. Phys. USSR 11, 23 (1947).
[10] A. Arima, T. Otsuka, F. Iachello and I. Talmi, Phys. Lett. 66B, 205
(1977); T. Otsuka, A. Arima, F. Iachello and I. Talmi, Phys. Lett.
76B, 139 (1978).
[11] For a review, see F. Iachello and P. van Isacker, The Interacting
Boson-Fermion Model, (Cambridge University Press, 1991).
[12] F. Iachello and I. Talmi, Rev. Mod. Phys. 59, 339 (1987).
[13] F. Iachello, E. Lipparini and A. Ventura, Lecture Notes in Physics
404, 318 (Springer, Berlin, 1992).
18
Kaon Condensation in Dense Matter
Gerald E. Brown
Department of Physics
State University of New York at Stony Brook
Stony Brook, New York 11794
USA
Abstract
The K~-meson of mass m^ = 494 MeV is similar to an exciton, consisting
of a strange quark, as particle, and u antiquark, as hole, bound together to
make the kaon. In dense nuclear matter the kaon feels an attractive mean
field from the nucleons which lowers its energy appreciably. As soon as its
energy is brought down to ~ half of its rest mass, it becomes energetically
favorable in neutron stars to replace electrons by kaons, the neutron stars
becoming nuclear matter stars at higher density. The kaons form a zero
momentum Bose-Einstein condensate.
The new equation of state of dense matter, with the inclusion of kaon
condensation, is substantially softer than conventional ones, meaning the
maximum mass of compact objects formed in the collapse of large stars is
only ~ 1.5 M o . With this equation of state, black holes are easy to form
and it is estimated that there are ~ 109 of them in our galaxy. In this sense,
a large number of black holes is the "smoking gun" for kaon condensation.
1 Introduction
438
Kaon Condensation in Dense Matter 439
The K~-meson has a mass of
mK = 494 MeV . (1)
The kaon is a boson, which will be important for our condensation
scenario. It is composed of a strange quark and a nonstrange u antiquark;
i.e., \K~) = |us). Whereas we have heard, throughout this conference,
about the breaking of symmetry, in order to establish a Bose-Einstein
condensate, we should realize that from the viewpoint of QCD we live
in a world of broken symmetry, broken chiral symmetry. One of the
consequences of this is that the negative energy sea of quarks can be
considered to be filled, at least up to some cutoff momentum at which
asymptotic freedom takes over. The filled negative energy sea of quarks
plays the role of the valence band in CU2O, which is much discussed in
this volume (see the articles by Wolfe et al. and Mysyrowicz.) The u is
then a hole in the valence band of up quarks, and the s is a particle in
the conduction band of strange quarks. The K~ can, then, be considered
to be an exciton, with the s-particle somewhat heavier than the u-hole.
We shall work at temperatures very small compared with m^, a
few MeV in magnitude, and will assume for our purposes here that
T = 0. Temperature dependent effects have not yet been calculated for
kaon condensation. Our Bose-Einstein condensate is then quite trivial, a
condensate at zero temperature of K~-mesons in the state of zero mo-
mentum. I will try to convince you that such a condensate is most likely
formed in the collapse of a large star, and that kaon condensates are
integral components of the compact objects, often referred to as neutron
stars, formed in such collapses. Obviously, the question to be answered
is how the interactions compensate the mass (1) of the kaon. It must be
energetically favorable to introduce kaons, before they will appear.
We will develop the theoretical framework in the next section, or at
least, discuss it. Before doing this, let us review what has been deduced
empirically about the K~-nucleus interaction, using this knowledge to
build the K~-nuclear matter interaction. The K~-nucleus interaction has
been obtained recently by Friedman, Gal and Batty [1], who described
the kaonic atoms; i.e., the atoms formed by a K~-meson and the nucleus,
for nuclei across the periodic table. In the case of thef X-nucleon
interaction, there is the well-known A(1405) resonance, which lies below
the KN continuum, and thus would be a bound state if its decay through
the (TC,A) or (TC,Z) channels would be turned off [2, 3]. As it is, the
t The notation is that the K, the antikaon, has two components, K~ and K°. Similarly
K = (K+,K°).
440 G. E. Brown
A(1405) is a bound state in the continuum. Because of this bound state
below the X~-nucleon scattering threshold, the low-energy K~-nucleon
scattering amplitudes are repulsive and the optical model potential (mean
field potential)
2mKVmf = -4nbp , (2)
where b is the scattering amplitude, is repulsive. (Kinematical factors are
suppressed here. They go out in the end.)
However, since the K and nucleon form a bound state, albeit in
the continuum, the basic interaction between K and nucleon must be
attractive.
Friedman et al. [1] argue that as the nuclear density is increased, the
K~-nucleus interaction must turn attractive. From fitting kaonic atoms
across the periodic table, they deduce that the K~ in the middle of the
nickel nucleus (i.e., roughly at nuclear matter density po) experiences an
attraction of
Vmf(po) * -200 MeV . (3)
The precise value of this number is not so certain, because the analysis
is difficult due to the fact that the kaon can disappear through reactions
such as
K~ + N -> 2T + n (4)
and others; i.e., there is a large imaginary part to the potential. Let
us none the less accept that there is a large attraction of the order of
that given in (3). The simplest possible generalization of (3) to higher
densities is
Vmf(p) = -200(p/p 0 ) MeV. (5)
The energy of a K~-meson in dense nuclear matter then will be
eK- = 494 MeV - 200(p/p0) MeV . (6)
In the collapse of large stars, due to the presence of electrons a large
negative chemical potential \i- is established in the core of the collapsing
star. An example of such a star was the progenitor of Supernova 1987A,
which was known from observations of its luminosity and type, to be of
mass
Mstar = 18 + 2 M O , (7)
measured in units of the mass of our sun M o . In collapse calculations, the
Kaon Condensation in Dense Matter 441
core of the star which became the compact object, often called neutron
star, was estimated in many calculations to be
More important for us, the density reached in the center of the core is
generally found to be several times nuclear matter density po, the precise
value depending upon the equation of state employed.
During the collapse of the star and the supernova explosion, the
neutrinos formed in the capture reaction
e~ + p <-> n + v (9)
fie + Up = Hn + Mv • (10)
However, later, as neutrinos leave, the /xv can be left out of this equation
and we obtain
lie = lin-fip. (11)
e ^ K ~ + v, (14)
the neutrinos leaving the star.f We will give results of detailed calcula-
tions later.
The energy is lowest, in a mean-field description, if the kaons so formed
t The formation of kaon condensation will happen after the era of neutrino trapping, so
that neutrinos can freely leave the star.
442 G. E. Brown
go into a state of zero momentum. The kaon condensate can then be
described by a classical wave function (see (15) below).
Calculations to date of kaon condensation have been carried out at
zero temperature. However, temperatures as high as 75 MeV are reached
in the collapse of stars [4]. Characteristic thermal energies are eth ~ TTT.
Thus, eth is of the same order as the electron chemical potential /ie ~
225 MeV at condensation. Consequently, the thermal energy may hinder
Bose-Einstein condensation, as it seems to hinder the condensations of
spin-polarized hydrogen or excitons discussed in this volume. In the
case of excitons in CU2O discussed by Wolfe et al, modification of the
crystal symmetry by applied stress seems to give the excitons a new way
to eliminate thermal energy and entropy. Such a way is present in a
star. Neutrinos will carry off energy quickly, reducing the entropy to that
needed for Bose-Einstein condensation. But the precise time scale for all
of this still needs to be worked out.
9
Fig. 1. Ginzberg-Landau energy (the temperature T is taken to be zero) for
p > pc, where pc is the critical density for the phase transition.
have a pole at co = /*_. Here n(co) is the kaon self energy. In our model
it includes only s-wave interactions. Physically, the reason that co = ja-
is that given in the last section; below this energy it becomes favorable
for electrons to change into kaons.
We have investigated only the second order phase transition, which
takes place at pc = 3.22po, where
for p > pc. Here we have rescaled 8 so that the coefficient of the 0 4-term
is \. The coefficient a(p) has the form
fl(p) = a ( p - p c ) (21)
444 G. E. Brown
for a second order phase transition. The a comes out of our GL-
Lagrangian. Minimizing the energy with respect to 6 gives
— v — ±*\*\yjv ,- Zu , \^Z)
E = ~oc2(p-pc)2 . (24)
From this one can see that the gain in energy is very rapid above p = pc.
In Table 1 we show the results of calculations [7] for the scenario
outlined in the previous section. For completeness, we list refs. [8, 9, 10,
11, 12] as a chronological development of the development of the input
used in the GL equations.
Table 1 describes the composition of neutron stars at densities p > pc.
One sees that the proton fraction x rises rapidly with density towards 0.5
and that the electron fraction xe decreases. The dense part of the core is
no longer a neutron star, but nuclear matter, or a "nucleon" star. Thus,
the kaon condensation scenario has replaced neutron by nucleon stars.
This should have many major implications for their structure, which are
presently being worked out.
The tendency to go from neutron matter towards nuclear matter is
easily understood. Once the electrons can be replaced by kaons, the
high energies of the former (fermions) need not be introduced in order
to neutralize the charge on the protons. Other things being equal, the
strong neutron-proton attraction favors equal numbers of protons and
neutrons. In this way the nuclear symmetry energy can be brought lowest.
3 Lots of Black Holes are the Smoking Gun for Kaon Condensation
We would now like to discuss the effect of kaon condensation on the
structure of compact objects formed following the collapse of large stars.
In the last section, we saw that kaon condensation substantially lowered
the energy of the dense matter. It also greatly softens the pressure, so
that the maximum mass of the compact objects is lower than is usually
found in the literature. Indeed, Brown [13] found Mmax = 15 M o , just
about the same as (8), the estimated mass of the core formed in the
Kaon Condensation in Dense Matter 445
u e Ae A* X xK xe
3.22 0.0 0.0 224.0 0.163 0.000 0.096 0.067
3.72 24.1 -5.2 165.6 0.317 0.267 0.034 0.016
4.22 27.1 -15.8 127.4 0.385 0.368 0.013 0.003
4.72 27.4 -28.4 100.8 0.421 0.415 0.006 0.000
5.22 26.8 -41.9 81.3 0.444 0.441 0.003 0.000
5.72 26.0 -55.9 66.7 0.458 0.457 0.001 0.000
6.22 25.0 -70.1 55.5 0.468 0.467 0.001 0.000
6.72 24.1 -84.7 46.7 0.475 0.475 0.000 0.000
7.22 23.1 -99.4 39.7 0.480 0.480 0.000 0.000
7.72 22.3 -114.3 34.0 0.484 0.484 0.000 0.000
8.22 21.4 -129.5 29.4 0.487 0.487 0.000 0.000
8.72 20.6 -144.9 25.6 0.489 0.489 0.000 0.000
9.22 19.9 -160.5 22.4 0.491 0.491 0.000 0.000
explosion of SN 1987A. It has been suggested [14] that this core went
into a black hole.
Arguments have been given [15] that the compression modulus of
nuclear matter is somewhat smaller than the conventional value [16]
Ko = 210 + 30 MeV. It was found [13], however, that Ko had to be raised
to 190 MeV, the lower region of conventional values, in order to stabilize
a 1.50 M o nucleon star. Figure 2 shows the known measured masses
of neutron stars. It is quite striking that all of the accurately measured
masses in binaries lie below 1.5 M o , the largest being the 1.44 M o pulsar
in PSR 1913+16. The mass of Vela X-l lies somewhat higher, but the
relatively large error bars reach down to ~ 1.55 M o .
In the formation of binary pulsars, after the first large star of a
binary explodes in order to form a compact object, the latter and the
remaining large star share a period of common envelope. During this
period, accretion can be rather large, ~ 10~3 M o per year [18]. Given
this large possible accretion, it would seem strange if compact objects
larger than 1.44 MG would not be seen in binaries, were they to exist.
Of course, 1.44 M o is just the Chandrasekhar mass for an N = Z star,
so that the collapse of the cores of large stars begins only when they
446 G. E. Brown
4U 1538-52
Ccn X 3 i •
HerX-1 •—•-
SMCX-1 i •
Vela X - l I » i
1 2
Neutron star mass (Mo)
are unstable, with masses somewhat less than 1.44 M 0 because of some
electron capture. However, with accretion following the collapse, the
compact cores formed from large stars are estimated to be of the order
1.4 M o . Thus, it is true that compact objects tend to be formed with
such a mass. Our argument is that accretion from a companion star
should produce, in some cases, a larger mass. Since these are not seen,
this argues in favor of the maximum mass being not much larger than
1.44 M o .
We next argue that the collapse scenario of the core of a large star
is radically changed at high densities because of the kaon condensate.
Let us first review the conventional collapse scenario. As can be seen
from Fig. 3, the maximum masses of neutron stars at short times are
less than the longtime cold maximum mass, even though the trapped
neutrinos give considerable pressure. The point is that nuclear matter is
substantially softer than neutron matter, and the former is turned into
the latter as the trapped neutrinos leave, allowing electron capture. The
conversion of protons to neutrons increases the pressure more than it is
decreased by the loss of neutrinos. Thus, if the mass of the compact object
exceeds 1.5 M o at early times (or any time), it will immediately collapse
into a black hole. Inclusion of thermal pressures, which are presently
being calculated, may raise the solid, short-time curve, slightly above
the dashed, long-time one, so that there would be a small "window"
Kaon Condensation in Dense Matter 447
of masses for a star to exist for some seconds, and then collapse into a
black hole as it cooled. We shall next discuss what happens in connection
with the much larger window, found in the case of the equation of state
including kaon condensation.
In the case of kaon condensation, the original nuclear matter does not
change into neutron matter. Indeed, at the higher densities, the fraction
of protons goes to one half. The early time maximum masses, as a
function of central density, are shown by the solid line in Fig. 4.
As can be seen from Fig. 4, the compact object can be stabilized,
during the time that the neutrinos are trapped, and then collapse into a
black hole. This means that the star can explode, returning matter to the
galaxy, and then go into a black hole [20]. Such black holes will be of
relatively light mass, not much larger than 1.5 M o . In general, observers
have looked for invisible centers of gravitational attraction, heavier than
448 G. E. Brown
2.0 1
1 • 1 1 ' 1 ' 1 ' 1
1.5 - 1
II
ll
if
1
II
1.0 - -
0.5 - -
0.0 1 1 . | . 1 . 1 . 1 . 1
0.0 1.0 2.0 3.0
Fig. 4. As in Fig. 3 the solid line shows the short-time maximum nucleon star
mass, as function of central density, with an assumed lepton fraction of Yf —
0.4, consisting of electrons and trapped neutrinos. The dashed line shows the
long-time cold mass, after the neutrinos have left.
^=4±1.3. (25)
AZ
If all stable stars of mass up to ~ 100 M o were to explode, returning
matter to the galaxy, this ratio would lie between 1 and 2. Helium is
produced chiefly by relatively light stars, metals by heavy stars, so that
cutting off the production by the heavy stars going directly into black
holes without nucleosynthesis increases the AY/AZ. Maeder [22] found,
using the standard initial mass function for stars, that Pagel's measured
AY/AZ was best reproduced by a cutoff of nucleosynthesis at a main
sequence stellar mass of ~ 22.5 M o . There is considerable uncertainty
in the initial mass function, as noted by Maeder [22], so that this limit
could easily be ~ 30 M o or even higher.
The Brown-Bethe estimate [20] of 30 M o as the cutoff for stars to
drop directly into black holes without nucleosynthesis and the scenario
discussed above that a large range of stars below this mass can first
accomplish nucleosynthesis and then collapse into (light mass) black
holes, extends the estimated range of main sequence masses for which
stars go into black holes down to ~ 18 M o [20]. Interestingly, this
possibly includes Supernova 1987A, with progenitor mass of 18 + 2 M o .
Indeed, arguments have been put forward [14] that SN 1987A probably
went into a black hole.
A good candidate for a similar situation is the supernova remnant Cas
A, where the supernova explosion — unobserved at the time — took
place in the late seventeenth century. From element abundances in the
450 G. E. Brown
ejecta it has been argued [24] that the progenitor was an ~ 20 M o star, of
roughly equal mass to the progenitor of SN 1987A. Although observers
have looked with high accuracy in several different ways at the center of
the supernova remnant, they have been unable to find a compact object.
The Brown-Bethe scenario suggests that about as many stars end up
in black holes as nucleon stars. They estimate that there are ~ 500
million heavy mass black holes and a comparable number of light mass
black holes in the Galaxy. In fact, compact remnants are seen in only
~ 20 of the roughly 150 supernova remnants in the Galaxy [25]. It may
turn out that a black hole is the more likely fate than a nucleon star for
the compact core of a large star.
I would like to thank M. Prakash for pointing out the large "window"
for nucleosynthesis, with later collapse into a black hole, in the case of
strangeness condensates, and for providing me with Figs. 3 and 4. I am
grateful to Vesteinn Thorsson for providing me with Table 1 and both to
him and to Mannque Rho for their collaboration on the theory of kaon
condensation. I am grateful to Hans Bethe, with whom the scenario for
black hole formation of ref. [20] was worked out.
This work was supported by the US Department of Energy Grant No.
DE-FG02-88ER40388.
References
[1] E. Friedman, A. Gal and C. J. Batty, Phys. Lett. B 308, 6 (1993).
[2] P. B. Siegel and W. Weise, Phys. Rev. C 38, 221 (1988).
[3] A. Muller-Groeling, K. Holinde and J. Speth, Nucl. Phys. A 513,
557 (1990).
[4] J. Cooperstein, in First symposium on nuclear physics in the universe,
ORNL, TN, 2^26 Sept. (1992), M. Guidry, ed. (Adam Hilger), in
press.
[5] G. E. Brown, K. Kubodera and M. Rho, Phys. Lett. B 192, 273
(1987).
[6] G. Baym and D. K. Campbell, Mesons in Nuclei III, M. Rho and
D. H. Wilkinson, eds. (North-Holland, Amsterdam, 1979), p. 1031.
[7] V. Thorsson, private communication.
[8] D. B. Kaplan and A. E. Nelson, Phys. Lett. B 175, 57 (1986).
[9] H. D. Politzer and M. B. Wise, Phys. Lett. B 274, 156 (1991).
[10] G. E. Brown, K. Kubodera, M. Rho and V. Thorsson, Phys. Lett.
B 291, 355 (1992).
Kaon Condensation in Dense Matter 451
[11] G. E. Brown, C.-H. Lee, M. Rho and V. Thorsson, Nucl. Phys.
A 567, 937 (1994).
[12] G. E. Brown, Nucl. Phys. A, to be published.
[13] G. E. Brown, in First symposium on nuclear physics in the universe,
ORNL, TN, 24^26 Sept. (1992), M. Guidry, ed. (Adam Hilger), in
press.
[14] G. E. Brown, S. W. Bruenn and J. C. Wheeler, Comm. Astrophys.
16, 153 (1992).
[15] G. E. Brown, Nature 336, 519 (1988).
[16] J. P. Blaizot, D. Gogny and B. Grammaticos, Nucl. Phys. A 265,
315 (1976).
[17] S. E. Thorsett, Z. Arzoumanian, M. M. McKinnon and J. H. Taylor,
Astrophys. Journ. Lett, 405, L29 (1993).
[18] R. A. Chevalier, Astrophys. Journ. 346, 847 (1989) ; R. A. Chevalier,
Astrophys. Journ. 411, L33 (1993).
[19] M. Prakash, private communication. See also V. Thorsson, M.
Prakash and J. M. Lattimer, Nucl. Phys. A 572, 693 (1994).
[20] G. E. Brown and H. A. Bethe, Astrophys. Journ. 423, 659 (1994).
[21] S. E. Woosley and T. A. Weaver, Ann. Rev. Astron. and Astrophys.
24, 205 (1986); J. R. Wilson, R. Mayle, S. E. Woosley and T. A.
Weaver, Proc. 12th Texas Rel. Astrophys. Symp., Jerusalem, Israel,
16-20 Dec, (1984).
[22] A. Maeder, Astron. Astrophys. 264, 105 (1992).
[23] B. E. J. Pagel, E. A. Simonson, E. A. Terlevich and M. G. Edmunds,
MNRAS 255, 325 (1992).
[24] R. A. Chevalier and R. P. Kirshner, Astrophys. Journ. 233, 154
(1979).
[25] S. Kulkarni, private communication.
19
Broken Gauge Symmetry in a Bose
Condensate
A. J. Leggett
Department of Physics
University of Illinois at Urbana-Champaign
Urbana, IL 61801
USA
Abstract
452
Broken Gauge Symmetry in a Bose Condensate 453
takes the thermodynamic limit in a similar way, the quantity rj = < \p >
("order parameter") becomes finite (in this case proportional to N1^2)
and its phase q> thus becomes well defined. The Bose condensate is thus
said to possess "spontaneously broken gauge symmetry", just as the fer-
romagnet below its Curie temperature possesses "spontaneously broken
rotational symmetry".
The aim of the present paper is to examine in detail the concept of
spontaneously broken gauge symmetry and hence, by implication, the
degree of validity of the analogy between the Bose condensate and the
ferromagnet. To some extent, it is a continuation of the discussion
begun in Ref. [2], to which I shall sometimes refer for technical details
for which there is no space here. It should be said at once that, at
least as regards the vast majority of the experiments apparently feasible
today, nothing of any great importance hinges on the outcome of the
discussion, and indeed the points at issue might be regarded as in some
sense "theological"; my reason for pursuing them nevertheless is simply
that I find existing treatments of the concept of broken gauge symmetry
somewhat unsatisfactory. However, as a by-product we shall find that
some experiments are suggested which are not obviously beyond existing
capabilities.
I wish to make three remarks before I start on the main theme.
First, I shall feel free in the following to use examples, and thought-
experiments, which involve not only "Bose condensation" in the strict
sense but also the "pseudo-Bose condensation" of Cooper pairs in Fermi
systems. In the latter system, the quantity which plays the role of the
order parameter < xp > in a Bose system is related to the so-called
"two-particle anomalous average" defined by
M r y ) MV>«(OVVKO>. (1)
For example, in a simple BCS superconductor the natural choice for the
order parameter is the quantity F^(r9r), and all considerations regarding
the phase, etc., then go forward just as in the Bose case.t In the case of
Fermi superfluids such as 3He with more complicated pairing structures,
there may be more than one order parameter (as in the spin-1/2 Bose
system formed by spin-polarized atomic hydrogen, see Ref. [2]) and their
relative phase then becomes of interest: see Ref. [2], and below.
Secondly, for the purposes of the present paper the relevant order
parameter will always be assumed to be constant in space, except when
t Except for a factor of 2 in the commutation relations, which is irrelevant to the
considerations of this paper and which I will generally ignore.
454 A. J. Leggett
there is a discrete discontinuity such as that across a Josephson junc-
tion. To be sure, the question of how far the concept of spontaneously
broken symmetry remains valid in the presence of substantial spatial
inhomogeneity is an important one, and may well be connected with
some of the intriguing questions raised by Kagan [3], but to keep the
discussion focused I shall not attempt to attack it here. However, one
may hope that an improved understanding of the conditions for, and
implications of, broken gauge symmetry in the simpler cases discussed
here may eventually shed some light on these more practically important
questions.
Thirdly, let us try to tighten up the Bose-ferromagnet analogy a little
by focusing specifically on two variables of the ferromagnet, namely the
z-component of total spin Sz and the angle 9 = tan~1(5>,/5x) made by the
component of spin in the xy plane with some arbitrary reference axis. We
could, in fact, at this point drop the requirement that the Hamiltonian of
the ferromagnet has the complete 0(3) symmetry and demand only that
it be invariant under rotation around the z-axis. Under these conditions
the formal analogy with the Bose system is complete: to make it explicit,
we define the operator S+ = Sx + iSy so that we have the commutation
relation
S+, (2)
in complete analogy with the relation for the Bose case
[N,xp+]=ip+, (3)
where N is the total number operator. Thus, Sz is the analog of N and
(—)6 of the "phase" cp of xp; Sz generates "rotations" of 6 just as N
does of cp. We shall tentatively treat 6 and cp as operators satisfying
the commutation relations [5Z, 6] = j, [AT, cp] = —i. Although there are,
of course, well known difficulties [4] in doing so when the number
of particles involved is small, these difficulties appear to be irrelevant
for systems of typical "laboratory" size (as is shown, inter alia, by the
success of theories in which the phase is treated as a quantum-mechanical
operator in predicting the rate of "macroscopic quantum tunneling" in
Josephson systems, see e.g. Ref. [5]), and in any case do not appear to
affect the questions that I shall consider in this paper. The meaning of
the relative phase as an operator is clarified explicitly below. Of course,
it is conceivable that in the comparison of the role of cp with that of 0,
a crucial role is played by the possibility that the ferromagnet has the
full 0(3) symmetry and not just the 1/(1) one corresponding to rotation
Broken Gauge Symmetry in a Bose Condensate 455
around the z-axis; we return briefly to the possible relevance of this
circumstance at the end of the paper.
Let us first review briefly some of the considerations discussed in more
detail in Ref. [2]. We first note that for an isolated system with a fixed
number JV of particles it is possible to define the "order parameter" rj(r)
in an alternative way, namely
where cp is explicitly the phase of xo- Since the latter has no physical
meaning, it is clear that neither does the phase of rj, as we should of
course infer, for definite N, from (3). Thus it is clear that to make sense of
the concept of the "phase" of the order parameter, we must first discuss
the conditions for defining relative phase, and then ask whether we can
in some sense define an "absolute" phase by means, for example, of a
universal "phase standard". Note at this point that we do not normally
regard a similar problem as arising for the ferromagnet, because we are
used to the idea that Sz need not be well defined (thus allowing 9 to
have a definite value), whereas the notion that N is not definite is more
difficult to accept. Whether or not this difference of perception is really
well based is a question we shall have to return to later.
Let us then consider, as in Ref. [2] , a system which has available to
it two states xu li i n t o which Bose condensation may take place. An
obvious example is two bulk superconductors (or superfluids) joined by
a Josephson junction; in this case the states xi and xi would of course
simply correspond to localization on one side of the junction or the
other. However, it is more convenient to discuss first a case which is
closer to that envisioned in section 3 of Ref. [2], namely the Cooper
pairs in superfluid 3He-A. If we ignore the orbital degrees of freedom,
and moreover suppose that the so-called d-vector (for notation, see e.g.
Ref. [6]) lies in the xy-plane, then the two states x\ a n d Xi m question
correspond to pairs formed with total spin ±1, respectively, along the
z-axis: we denote these by ||f > and \[{> respectively. The most general
form of wave function in which all Cooper pairs are "Bose-condensed"
456 A. J. Leggett
into a single wave function may thus be written schematically (a, b real)f
(TV being the total number of 3He atoms). In the theory of magnetic
resonance in superfluid 3He [7], the ground state wave function is taken
to be of the form (6) with a = b and the relative phase Acp of the up-
and down-spin pairs chosen so as to minimize the dipole energy4
What is the variable canonically conjugate to Acpl It is nothing but the
difference AN of the total number of up and down spins in the system
(actually, of course, this is the total z-component of spin within a factor,
but I avoid the notation Sz so as to reserve this for the ferromagnetic
analogy). That AN and Acp are indeed canonically conjugate variables
(in the limit N -• oo) may be seen in a number of different ways, e.g.
by going back to the microscopic definition (1) of the order parameter
(see Ref. [7], section 2). For our purposes the most instructive derivation
probably consists in considering the class of many-body wave functions
(a subspace of the complete many-body space, of course) defined by
¥ = f c(AcpmAcp)dA(p, (7)
(9)
f Needless to say, even if the answer were to turn out to be yes, this is unlikely in the
extreme to be a practical proposition, inter alia because of the extreme difficulty of
controlling the chemical potentials acting on the different buckets sufficiently accurately
to prevent an unknown drift of their relative phase.
460 A. J. Leggett
Suppose further that the same is true of S and 2 (for the precise meaning
of this statement, see below). Is it then true that 1 and 2 have a definite
relative phase? Note that this question is experimentally answerable via
an ensemble of experiments in which 1 and 2 are placed in contact and
the "instantaneous" Josephson current between them is measured. If the
relative phase is definite, the same current will flow on all occasions,!
while if it is not true, the results will be randomly distributed.
It is, however, crucial to distinguish two different versions of this
thought-experiment. In version (A), S is placed in contact with 1 and 2
simultaneously and the whole system is allowed to come to equilibrium;
all contacts are then broken, and 1 is subsequently placed in contact with
2 without the mediation of S. In version (B), 1 is first placed in contact
with S, with 2 absent, and sufficient time is allowed for equilibrium
to be established. The 1-S contact is then broken, and subsequently
contact between S and 2 is made and time allowed for the attainment of
equilibrium. Finally this contact is also broken, and eventually 1 and 2
are put in contact as in version (A).
In view of space limitations, I shall give here only the results of the
analysis of these two thought-experiments; the details of the calculations,
which are not entirely trivial, will be presented elsewhere [10]. Case (A) is
straightforward: at the end of the day, a definite phase relation is indeed
established (subject, of course, to all the necessary caveats concerning
time-scales, etc., see Ref. [2]) between the superfluids 1 and 2, so that
on joining them a predictable Josephson current is obtained. Case (B)
is a little more tricky: one might intuitively suspect that the phase
relation established at the first stage between 1 and S would be "upset"
in the subsequent attainment of Josephson equilibrium between S and
2, and this indeed turns out to be the case, although, surprisingly, a
proper demonstration of this result requires explicit consideration of
the "environment" of the system (e.g., for the superconducting case,
the radiation field). Indeed, given a certain very plausible, though not
rigorously proved, assumption [10] concerning the dependence on the
total particle number involved of the final state of the environment, one
can demonstrate that no definite phase relation is established at the end
of the day between systems 1 and 2. Hence a "standard of superfluid
phase" is possible only in sense (A), not in sense (B).
We must finally return to the ferromagnetic analogy. Let us try, as
t In order to obtain a finite Josephson current, one could apply a known chemical potential
difference between 1 and 2 for a known time (so as to make Acpn nonzero before placing
them in contact).
Broken Gauge Symmetry in a Bose Condensate 461
a thought-experiment, repeating the arguments made in this paper with
the "phase of a superfluid" replaced by the "direction of spin (in the xy-
plane) of a ferromagnet". Just as we have implicitly assumed throughout
that the total number of the relevant particles (electrons, He atoms, etc.)
in the universe is concerned, let us for the sake of the argument make the
analogous assumption of conservation of the total z-component of spin.f
We would then have to conclude that, strictly speaking, the absolute
direction of magnetization is meaningless and that we should only talk
about relative directions of spin of different magnets. That most of us are
happy in everyday life with the idea that for all practical purposes we can
talk about "absolute" directions would then seem to imply a tacit belief in
the possibility of "standardization". For the case of the superfluid phase,
we have argued that such standardization is possible only in sense (A),
which requires prima facie very special and artificial conditions: what, if
any, is the salient difference in the case of magnetization?
I believe that the crucial difference between the two cases is to be
sought in the very different role of what we have vaguely called the
"environment". It turns out to be implicit in our arguments about phase
standardization for a superfluid that the "environment" (radiation field,
phonons, normal electrons or whatever) either cannot exchange particles
with the "system" at all, or, even if it can, cannot exist (or is not typically
found to exist) in the coherent superposition of states of different particle
number (in the sense, explored above, in which a superfluid system does).
That this is so is essentially just the statement that most of the world
is not superfluid (let alone a superfluid state of the relevant particles) !J
On the other hand, just about any "environment" can not only exchange
angular momentum with the system, but has no difficulty in existing
in a coherent superposition of states of different Sz (or rather Jz: here
the exchange of spin and orbital angular momentum does play a vital
role). More accurately, the universe can easily exist in a superposition
of states which differ by the relative angular momentum of "system"
and "environment". (That this is so is no doubt not unconnected with
the question of the more general SO(3) invariance.) In other words, the
typical situation as regards magnetization corresponds to "version A"
f Of course in real life spin and orbital angular momentum are not separately conserved,
but taking this into account only pushes the problem one step back.
$ To be sure, a superconductor may induce a degree of phase coherence in a normal metal
adjacent to it - the so-called proximity effect (see e.g. Ref. [11], section 7.3). However,
the effects die out exponentially with distance into the normal metal and thus cannot
play the necessary "global" role.
462 A. J. Leggett
of our thought-experiment rather than "version B", so the problem of
standardization is automatically overcome.
A final thought: Will "version A" ever be realized in a practical sense
for superfluid systems? This idea is not quite so preposterous as it
sounds. If the dream of room-temperature superconductivity is attained
(and perhaps even if it is not), it is not impossible that by the year 2100
there will be a network of superconducting transmission lines covering
the globe, with no normal "breaks". Then, indeed, it would be possible
in principle for groups, based say, in Tokyo and Washington, DC, to
"mate" two pieces of superconductor which had never been in direct
physical contact in such a way as to produce a predictable Josephson
current on first acquaintance! Although this state of affairs would of
course in no way alter any of the considerations present above, it might
perhaps modify our perception of their significance somewhat.
This work was supported by the National Science Foundation under
Grant No. DMR 92-14236.
References
[I] RW. Anderson, Basic Notions of Condensed Matter Physics (W.A.
Benjamin, Menlo Park, California, 1984).
[2] A.J. Leggett and F. Sols, Found. Phys. 21, 353 (1991).
[3] Yu. Kagan, this volume.
[4] P. Carruthers and M. M. Nieto. Phys. Rev. Lett. 14, 387 (1965).
[5] A. J. Leggett, J. Phys. Soc. Japan Supp. 26-3, 1986 (1987).
[6] A. J. Leggett, Rev. Mod. Phys. 47, 331(1975).
[7] A. J. Leggett, Ann. Phys. 85, 11(1974).
[8] A. J. Leggett, in Chance and Matter, J. Souletie, J. Vannimenus and
R. Stora, eds. (North-Holland, Amsterdam, 1987).
[9] R. Hegstrom and F. Sols, submitted to Phys. Rev. A.
[10] A. J. Leggett, Found. Phys., to be published.
[II] P. G. de Gennes, Superconductivity of Metals and Alloys (W.A.
Benjamin, N.Y, 1966).
Part two
Brief Reports
20
BEC in Ultra-cold Cesium: Collisional
Constraints
E. Tiesinga, A. J. Moerdijk, B. J. Verhaar, and H. T. C. Stoof
Department of Physics, Eindhoven University of Technology,
P.O. Box 513
5600 MB Eindhoven
The Netherlands
Abstract
465
466 E. Tiesinga, A. J. Moerdijk, B. J. Verhaar, and H. T. C. Stoof
where n is the density and A is the temperature-dependent DeBroglie
wavelength. The first system considered was a gas of ultracold hydrogen
[1] stored either in a gas cell with superfluid helium-lined walls or in a
wall-free magnetic trap. Although enormous progress has been made,
the required density/temperature combination has not been achieved.
Since the advent of laser cooling [2] other species have become avail-
able as candidates for observing Bose-Einstein condensation. In a few
milliseconds the (mostly alkali-metal) atoms are cooled from room tem-
perature to temperatures in the micro kelvin regime and after that stored
magnetically. One of the first experiments to magnetically trap heavy
alkali-metal atoms used 133Cs [3]. In this case the attempts so far reached
a density/temperature combination of 2 x 1010 cm"3 and 1//K. All exper-
iments aimed at Bose-Einstein condensation are hampered by collisions
between atoms. This limits the density and the lifetime of the sample to
a large extent. Since storage in a magnetic trap is restricted to low-field-
seeking hyperfine states of the electronic ground state, particle loss occurs
either by exothermic collisions releasing sufficient kinetic energy to leave
the trap or by transitions to high-field-seeking states. For sufficiently
high densities recombination is an important loss mechanism as well.
Although inelastic collisions limit the observable densities, the elastic
collisions are essential for the formation of the condensate [4]. This is
due to the fact that these collisions comprise the only process available
to transfer the atoms in the one-particle ground state. Furthermore, the
validity of the usual discussion of the Bose condensate in terms of a
Bogoliubov transformation [5] crucially depends on a positive value of
the scattering length, which is equivalent to a repulsive elastic collision
at low kinetic energies. For negative values the Bogoliubov theory breaks
down. This paper briefly discusses the trap-loss effects for atomic cesium
and focusses mainly on the resonant behavior of the scattering length
as a function of magnetic field. The latter promises to be an excellent
property for the study of the Bose condensate. An extensive discussion
of the density decay can be found in Ref. [6].
The starting point of the our discussion is the hyperfine level diagram
(Fig. 1) of the cesium ground-state (2Si/2) atom. If we consider the
magnetic field along the z-axis, the effective single-atom Hamiltonian
comprises hyperfine and Zeeman terms
where Se and S" are electron and nuclear spin, respectively. Magnetic
BEC in Ultra-cold Cesium: Collisional Constraints 467
traps generally operate with fields well below the critical value B c =
4ahf/(yeh) « 0.33 T, so that the total spin vector f = S e + Sn is still
conserved. Since the electron (nuclear) spin is 1/2 (7/2), the 16 hyperfine
states of cesium are conveniently labeled with |/,m/). The collision of
two such atoms is described by the effective two-body Hamiltonian [6],[7]
H = Z- (2)
containing a kinetic energy term with ft the reduced mass and two-body
interactions Vc and Vd. The central or exchange interaction Vc represents
an effective description of all Coulomb interactions between the electrons
and nuclei and depends only on the magnitude of the total electron spin
S = Sf + S|. It can be written as a sum of singlet and triplet terms
Vc = (3)
500
(0,6,-6)
(£,F,Mp)=(0,7,-6)
6 (0^8,-6)
i
t-l
J
O
CO
-500
0.00 0.05 0.10
References
[1] I.F. Silvera and J.T.M. Walraven, Phys. Rev. Lett. 44, 164 (1980);
J.M. Doyle, J.C. Sandberg, LA. Yu, C.L. Cesar, D. Kleppner, and TJ.
Greytak, Phys. Rev. Lett. 67, 603 (1991); O.J. Luiten, H.G.C. Werij,
I.D. Setija, M.W. Reynolds, T.W. Hijmans, and J.T.M. Walraven,
Phys. Rev. Lett. 70, 544 (1993).
[2] The special issues of J. Opt. Soc. Am. B devoted to laser cooling;
Vol. 2 (1985) and Vol. 6 (1989).
[3] C. Monroe, W. Swann, H. Robinson, and C. Wieman, Phys. Rev.Lett.
65, 1571 (1990); C. Monroe, E. Cornell, C. Sackett, C. Myatt,and C.
Wieman, Phys. Rev. Lett. 70, 414 (1993).
[4] H.T.C. Stoof, Phys. Rev. Lett. 66, 3148(1991); H.T.C. Stoof, Phys.
Rev. A 45, 8398 (1992); Yu. M. Kagan, and G.V. Shlyapnikov, Zh.
Eksp. Teor. Fiz. 101, 528 (1992); B.V.Svistunov, J. Moscow Phys.
Soc. 1, 373 (1991). Contributions of these authors on nucleation can
also be found in this volume.
BEC in Ultra-cold Cesium: Collisional Constraints 471
[5] A.L. Fetter and J.D. Walecka, Quantum Theory of Many-Particle
Systems, (McGraw-Hill, New York, 1971).
[6] E. Tiesinga, A.J. Moerdijk, BJ. Verhaar, and H.T.C. Stoof, Phys. Rev.
A 46, R1167 (1992); E. Tiesinga, BJ. Verhaar, and H.T.C. Stoof,
Phys. Rev. A, May (1993).
[7] H.T.C. Stoof, J.M.V.A. Koelman, and BJ. Verhaar, Phys. Rev. B 38,
4688 (1988).
[8] H. Weickenmeier, U. Diemer, W. Demtroder, and M. Broyer, Chem.
Phys. Lett. 124, 470 (1992); H. Weickenmeier, U. Diemer, M. Wahl,
M. Raab, W. Demtroder, and W. Muller, J. Chem. Phys. 82, 5354
(1985).
[9] M. Krauss and WJ. Stevens, J. Chem. Phys. 93, 4236(1990).
[10] P.S. Julienne, J. Mol. Spectrosc. 56, 270 (1975); S.R. Langhoff, J.
Chem. Phys. 61, 1708 (1974).
21
BEC and the Relaxation Explosion in
Magnetically Trapped Atomic Hydrogen
T. W. Hijmans, Yu. Kagan,* G. V. Shlyapnikov,* and J. T. M. Walraven
Van der Waals-Zeeman Laboratoriwn, Universiteit van Amsterdam,
Valckenierstraat 65/67
1018 XE Amsterdam
The Netherlands
Abstract
We predict and analyze non-trivial relaxational behavior of magnetically
trapped gases near the Bose condensation temperature Tc. Due to strong
compression of the condensate by the inhomogeneous trapping field, par-
ticularly at low densities, the relaxation rate shows a strong, almost jump
wise, increase below Tc. As a consequence the maximum fraction of con-
densate particles is limited to a few percent. This phenomenon can be called
a "relaxation explosion". We discuss its implications for the detectability
of BEC in atomic hydrogen.
472
BEC and the Relaxation Explosion in Trapped Atomic H 473
atoms (Hj) which are ejected from the trap. The approach to BEC
is often [6, 8] described in terms of trajectories in density-temperature
space. We shall see that in H, the increase in relaxation rate resulting from
the appearance of the condensate, the "relaxation explosion", prevents
one from penetrating deep into the BEC region and markedly alters these
trajectories.
For simplicity we consider an isotropic harmonic trapping potential of
the form
V(r) = fiBB0 + ™co 2r\ (1)
where r is the distance to the trap center, and co is the oscillation
frequency. The critical BEC temperature is expressed by the relation
Tc = 3.31ft2n2/3/m, where n is the density of the gas at the trap center
and m is the mass of the atom. The density profiles in external potentials
were analyzed by Goldman et al. [9], and by Huse and Siggia [10]. The
critical temperature can be expressed in terms of the total number of
particles N9 and the parameters of the trapping potential [11]. For the
potential given by Eq. (1) we have
Tc = (N/ g 3 (l)) 1 / 3 ^. (2)
Here g3(l) = 1.20, where g/ is a Bose integral given by g«K£) =
£^i(£7«0.
Well above Tc, the number of relaxation events per unit time is given
by
= 2a /drn 2 (r), (3)
where n(r) is the density distribution and a is the rate constant for
dipolar decay. For hydrogen a « 10~ 15 cm3/s [12, 13]. Below Tc we
should replace Eq. (3) by a more general expression [14, 15]:
where xp is the field operator of the atoms and < xp^xp^xpxp > is the local
two-particle correlator.
A detailed calculation [16] shows that, below Tc, the relaxation rate
can be expressed in terms of its value at Tc:
T AT
vr(T) = vr(Tc)(l + 7.5(-^) 3 ' 5 (—) 7 / 5 ) ; T < Tc (5)
Here, AT = Tc — T, and ncU represents the mean field interaction
474 T. W. Hijmans et al.
Fig. 1. Number of relaxation events per unit time for a fixed number of atoms,
normalized to the value at Tc. The solid line corresponds to nc = 1015 cm" 3 and
the dashed line to nc = 1013 cm" 3 .
v0 =
Clearly, as a result of the weak interaction, we have Vo < Ve. Hence,
even if No < N (and accordingly AT < T) the condensate can dominate
the relaxation rate. In Fig. 1 the value of v r (T)/v r (r c ) is given for H
atoms for nc = 1013cm"3 (Tc = 7JIK) and nc = 1015cm"3 (Tc = 150/iK).
To investigate the effect of the relaxation explosion, we use a simple
model for the kinetics of the cooling proces, in which we assume that
the cooling proceeds sufficiently slowly to consider the system as being
in thermal equilibrium. To obtain the trajectories in N — T space, we
start from the expressions for the internal energy of the trapped gas at
BEC and the Relaxation Explosion in Trapped Atomic H 475
temperatures above and below Tc (here we neglect ncll):
;T>TC (7)
\~W)
t 1J7
T = 4U iT*T- (10)
1016
1014
10-5 10-4 10-3
Temperature (K)
Fig. 2. Cooling trajectories for Rv =0.1 and 0.05 (solid lines) plotted as N/Ve
vs T. The long-dashed curve is the BEC line. The short-dashed curves represent
the trajectories corresponding to a system obeying Boltzmann statistics.
References
[1] H.F. Hess, G. Kochanski, J.M. Doyle, N. Masuhara, D. Kleppner,
and T.J. Greytak, Phys. Rev. Lett. 59, 672 (1987).
[2] R. van Roijen, J.J. Berkhout, S. Jaakkola, and J.T.M. Walraven,
Phys. Rev. Lett. 61, 931 (1988).
[3] J.M. Doyle, J.C. Sandberg, LA. Yu, C.L. Cesar, D. Kleppner, and
T.J. Greytak, Phys. Rev. Lett. 67, 603 (1991).
[4] OJ. Luiten, H.G.C. Werij, I.D. Setija, M.W. Reynolds, T.W. Hijmans,
and J.T.M. Walraven, Phys. Rev. Lett. 70, 544 (1993).
[5] I.D. Setija, H.G.C. Werij, O.J.Luiten, M.W. Reynolds, T.W. Hijmans,
and J.T.M. Walraven, Phys. Rev. Lett. 70, 2257 (1993).
[6] H.F. Hess, Phys. Rev. B 34, 3476 (1986).
[7] Yu. Kagan and G.V. Shlyapnikov, Phys. Lett. A 130, 483 (1988).
[8] T.J. Tommila, Europhys. Lett. 57, 314 (1986).
[9] V.V. Goldman, I.F. Silvera, and A.J. Leggett, Phys. Rev. B 24, 2870
(1981).
BEC and the Relaxation Explosion in Trapped Atomic H All
[10] D.A. Huse and E.D. Siggia, J. Low. Temp. Phys. 46, 137 (1982).
[11] V. Bagnato, D.E. Pritchard, and D. Kleppner, Phys. Rev. A 35 4354
(1987).
[12] A. Lagendijk, I.F. Silvera, and B.J. Verhaar, Phys. Rev. B 33, 626
(1986).
[13] H.T.C. Stoof, J.M.V.A. Koelman, and B.J. Verhaar, Phys. Rev. B 38,
4688 (1988).
[14] Yu. Kagan, B.V. Svistunov, and G.V. Shlyapnikov, Pis'ma Zh. Eksp.
Teor. Fiz. 48, 54 (1988) [JETP. Lett. 48, 56 (1988).
[15] H.T.C. Stoof, A.M.L. Jansen, J.M.V.A. Koelman, and B.J. Verhaar,
Phys. Rev. A 39, 3157 (1989).
[16] T.W. Hijmans, Yu. Kagan, G.V. Shlyapnikov, and J.T.M. Walraven,
Phys. Rev. B 48, 12886 (1993).
22
Quest for Kosterlitz-Thouless Transition
in Two-Dimensional Atomic Hydrogen
A. Matsubara,f T. Arai, S. Hotta, J. S. Korhonen, T. Mizusaki,
and A. Hirai $
Department of Physics
Kyoto University
Kyoto 606-01
Japan
Abstract
We have cooled atomic hydrogen adsorbed on liquid helium to very low
temperatures: a surface temperature of SI mK was obtained with a density
of 2 • 1012 atoms/cm 2 . The two-dimensional hydrogen gas was cooled on
a cold spot, which was inside a large sample cell. When the cold spot is
much colder than the rest of the sample cell the surface density on it is
controlled by the hydrogen flux into the sample cell. We found that at least
98% of the recombination heat was carried away from the cold spot by
the excited H2 molecules. There was a heat input to the cold spot which
was independent of the presence of atomic hydrogen and which prevented
cooling of the adsorbed hydrogen below 87 mK. We attribute this heat input
to thermal conduction by ripplons.
1 Introduction
The experiments with polarized atomic hydrogen [1, 2] are usually made
in sample cells whose walls are covered with a helium film. The atomic
hydrogen is adsorbed on the He surface and forms a two-dimensional
(2D) gas, which is expected to undergo a Kosterlitz-Thouless transition
[3] to a superfluid state. The critical temperature of the transition is
TKT=nh2a/2kBm , (1)
where m is the mass of the hydrogen atom and a is the surface density.
If the adsorbed hydrogen is in thermal equilibrium with the bulk gas, it
is difficult to control the surface density and temperature independently:
t Permanent address : Department of Physics, Osaka-City University, Sumiyoshi-ku,
Osaka, 558 Japan,
t Deceased on 31 December, 1992.
478
Quest for K-T Transition in 2D Atomic Hydrogen 479
the adsorption energy ea is about 1 K and when temperature is lowered
much below I K the surface density tends to increase exponentially and
soon approaches the saturation density 2 • 1014 atoms/cm2. This leads to
rapid three-body recombination at the surface, which prevents cooling
below TKT-
Our method of cooling 2D-HJ, is the following: inside a large sample
cell we prepared a small cold spot. The cold spot was thermally isolated
from the rest of the sample cell, and its temperature, Tcs, was kept
much lower than that of the sample cell. At sufficiently low TQS the
recombination occurs mainly on the cold spot. Then the surface density
is nearly independent of temperature and is determined by the input flux
of hydrogen to the sample cell. We found that although the recombination
occurs mainly on the cold spot, most of the recombination heat is carried
away from the cold spot by the excited H2 molecules.
2 Experimental Apparatus
We prepared doubly polarized atomic hydrogen, HJ^r, as follows. The
atomic hydrogen was produced at low temperature by rf discharge. It
was precooled by a baffle before it was stored in a buffer volume, which
was connected to the sample cell through a flow impedance. The buffer
and the sample cell were in a magnetic field of 8 T. If theflowimpedance,
volume of the buffer and the flux of hydrogen to the buffer are selected
properly, the atomic hydrogen in the buffer volume is doubly polarized.
The sample cell was a cylinder with radius 17.5 mm and height 75 mm.
Its walls were covered with superfluid 4He film of saturation thickness.
The cold spot was a copper disk of radius 2.2 mm. It was located at the
upper part of the sample cell. The cold spot surface was partly covered
with a sintered silver piece of radius 1.4 mm and estimated surface area
of 0.1m2. The cold spot was thermally isolated from the rest of the
copper sample cell by a piece of Stycast with an outer radius of 8 mm.
The discharge cell and the baffle were anchored to a heat exchanger
of the dilution refrigerator at a temperature of 0.5 K. The common
temperature of the buffer volume and the sample cell was fixed to a
temperature T#, between 290 and 430 mK, by a temperature controller.
The cold spot was connected by a mesh of copper wires to the mixing
chamber of the dilution refrigerator and its temperature was controlled
at Tcs.
We measured two quantities: the bulk density n of the hydrogen in
the sample cell and the increase in the heat input to the cold spot which
480 A. Matsubara et al.
was due to the introduction of hydrogen. In order to measure changes in
the heat input to the cold spot we fixed the temperatures of the mixing
chamber and the cold spot separately by temperature controllers. Then
the change in the feedback power of the cold spot controller was equal
(but opposite) to the power Q due to the introduction of hydrogen.
The bulk density n was measured by making all the hydrogen in the
sample cell recombine on a bolometer. The density was calculated from
the total recombination heat, which was obtained from the increase in
the sample cell temperature. With this method we were able to measure
only densities larger than 1012cm~3 but the absolute value of n was
obtained reliably. For smaller densities n was calculated from the rise in
the bolometer temperature.
(2)
Here V is the sample cell volume' Acs is the area of the cold spot; and
</) is the flux of HJ^r to the sample cell; Ks and Ls are the two- and
three-body recombination rates on the cold spot; R!jf*9 KeJ* and If^
are the effective one-, two- and three-body recombination rates, which
include the contribution from the recombination on the hot part of the
sample cell wall and are defined through equations R^n = RS(A/V)GH,
Kebffn2 = Kbn2+KS(A/V)(J2H and Lebffn3 = Lbn3+LS(A/V)(T3H. The sample
cell area is A and GH = nAGxp(ea/kBTH) is the surface density on the hot
walls of the sample cell; X is the thermal de Broglie wavelength. In our ex-
periments, only the recombination constants Ls = 1.2 x 10~24 cm4 s"1 [4]
and Rs = 8.5 cm"1 are important. The determination of Rs is discussed in
section 4. For values of Ka and L&, see Ref. [4], and for Ks, see Ref. [5].
In Eq. (2), the thermal velocity of the atoms in the bulk gas is vb =
m, and VbtiAcs/4 is the number of atoms colliding with the cold
Quest for K-T Transition in 2D Atomic Hydrogen 481
T s (mK)
jioo_
10 20
1
1/Ts (K- )
Fig. 1. The calculated saturation values of the bulk density n in the sample cell
and the surface density a on the cold spot as a function of the inverse temperature
Tg"1 of the adsorbed hydrogen. In region (i) the cold spot does not affect the
bulk density; in regions (ii) and (iii) the recombination occurs mainly at the cold
spot.
spot surface in unit time. The sticking coefficient Sb = 0.33 T#/K [5, 6] is
the probability that a surface collision leads to the adsorption of the atom.
The mean residence time on the surface is TS = (4As/vsSs)exp(ea/kBTs),
where A, v and S must now be evaluated at the temperature of the
adsorbed hydrogen, denoted by Ts. Due to the thermal resistance
between the adsorbed hydrogen and the cold spot, Ts may not be the
same as Tcs-
The saturation values of n and <r, given by Eq. (2), are plotted in Fig. 1
as a function of Ts, at constant TH = 310mK and $ = 1.1 • lO^s" 1 .
There are three distinct temperature regions: (i) the high temperature,
(ii) the desorption limited, and (iii) the recombination limited regions. In
the high temperature region the recombination on the cold spot can be
neglected. In regions (ii) and (iii) the recombination occurs mainly on
the cold spot. Then a is given by the equation Acs(Kso2 + Lscr3) = <f> and
is almost independent of the cold spot temperature. A desired value of a
can be obtained by selecting an appropriate hydrogen flux <j). In region
(ii) the probability that the adsorbed atom desorbs before recombination
is large, while in region (iii) nearly all the atoms adsorbed on the cold
spot recombine before being desorbed.
482 A. Matsubara et a\.
T(mK)
1 190 80 60
10
TH=310mK
10 15 20
1/r (K-1)
Fig. 2. The saturated bulk density of hydrogen vs. the inverse cold spot temper-
ature TQI (filled circles) and the inverse temperature of the adsorbed hydrogen
Tjf1 (open circles). Ts is calculated from the measured heat input to the cold
spot and the thermal resistance between the ripplon and the phonon systems.
The dashed line is a guide for the eyes, and the solid curve is calculated from
Eq. (2).
T C S (mK)
100 80 60
1/Tcs
Fig. 3. The heat input to the cold spot caused by the atomic hydrogen as a
function of the inverse cold spot temperature. The circles show the measured
data; the solid line is the power carried as the kinetic energy of the hydrogen
atoms. Qrecomb is the total power released in the recombinations. Only a fraction
smaller than 1.4% of QreComb goes to the cold spot.
All these data were taken under conditions such that the recombina-
tion occurs mainly on the cold spot. The total recombination heat power
Qrecomb> which is calculated from the flux (j>9 is shown in Fig. 3 by the
upper dashed line. Comparison with Qsat reveals that less than 1.4% of
the total recombination heat goes to the cold spot. This implies that the
H2 molecule, which is left in a highly excited state in the recombination,
cannot penetrate into liquid helium, and the recombination heat is deliv-
ered evenly to the sample cell. A similar result was obtained by Vasilyev
et ah [8] They measured an upper limit of 10% for the recombination
energy which is deposited into helium at the recombination site. The
heat input in Fig. 3 seems to vary roughly proportional to the density
nsat in Fig. 2. A mechanism which would show such a dependence is
heat transfer from the hot bulk hydrogen gas to the cold surface. The
calculated power Qkm, which is released when the hot hydrogen atoms
thermalize after being adsorbed, is shown in Fig. 3 by the solid line. The
local heat flux
= 2kB(TH - T(r))nvcc/4, (3)
484 A. Matsubara et al.
is calculated by using the measured density n. Here a = (3/2)5 is
the thermal accommodation coefficient, measured by Helffrich et al.
[9] Outside the cold spot, the radial distribution of the temperature,
T(r), is determined by the thermal conductivity of Stycast and is not
changed much by a small q^m- We estimate Qkm by integrating qkm
over the cold spot and the Stycast area. Qkin follows Qsat down to the
lowest cold spot temperature. It thus appears that the fraction of the
recombination heat which goes to the cold spot was much smaller than
1.4%.
We have used Qkin in estimating the difference between the tempera-
tures of the cold spot and the adsorbed hydrogen. According to Reynolds
et al. [7], the bottleneck in the heat conduction from the adsorbed hy-
drogen to the cold spot is between the ripplon and the phonon systems
in the helium. We thus assume that the adsorbed hydrogen is at the same
temperature as the ripplons and that the temperature of the phonons in
helium equals the cold spot temperature. Using the result of Reynolds
et al. we calculate the temperature difference T$ — TQS which is caused
by Qkin being dumped to the ripplons on the cold spot. The open cir-
cles in Fig. 2 show the measured nsat as a function of the calculated
7$. Agreement with ncaic(Ts) is good at Ts > 90 mK but at lower
temperatures the measured Qsat is too small to explain the temperature
difference Ts — TCs- This difference should be attributed to some resid-
ual heat input which does not depend on the presence of the atomic
hydrogen.
The saturation of Ts can be explained by the ripplon heat conduction.
Using the conduction measured by Mantz et al. [10, 11], we obtain a heat
input of about 30 nW to the cold spot under the conditions of Figs. 2
and 3. At low temperatures the heat carried by the ripplons is thus
much larger than Qsat and determines the minimum 7$. The calculated
minimum Ts = 85 mK is in good agreement with the value of 87 mK
obtained from Fig. 2.
In order to clarify the effect of the ripplon heat conduction, we mea-
sured nsat as a function of TH with fixed TQS = 80 mK. Using Eq. (2) we
estimated Ts from nsat; TS(TH) is shown by the circles in Fig. 4. The lines
are calculated by using the results of Reynolds et al. The dashed line
is obtained by considering only the ripplon heat conduction, while the
lower solid line results if only Qkin is taken into account. Both ripplons
and Qkm contribute to the upper solid line. Clearly, the heat carried by
ripplons becomes important for low T#.
Quest for K-T Transition in 2D Atomic Hydrogen 485
300 400
TH(mK)
5 Summary
Our experiments demonstrate that two-dimensional atomic hydrogen gas
can be cooled on a small cold area much below the temperature of the
bulk gas. In particular, we have shown that the recombination heat was
not the limiting factor of the minimum temperature of the adsorbed
hydrogen: the minimum T$ was determined by the residual heat input
to the ripplon system of the He-film. This residual heat flux was due to
ripplon heat conduction. In future experiments we will try to cool the
two-dimensional hydrogen further by using thinner 4He films and 3He
or 3He-4He films.
Finally, we note that although we do not at the moment have a direct
signal from the adsorbed hydrogen we believe that the KT transition can
be detected even in the present set-up. This is due to a large decrease
in the surface recombination rate when the KT transition takes place
[12,13]. If the measurement is done in the desorption limited temperature
region (ii) in Fig. 1, the transition should be seen as a sudden increase in
n when Ts is lowered below TKT>
486 A. Matsubara et al.
References
[I] I.F. Silvera and J.T.M. Walraven in Progress in Low temperature
Physics, Vol. 10, edited by D.F. Brewer (North Holland, Amsterdam,
1986), p. 139.
[2] I.F. Silvera and M. Reynolds, J. Low Temp. Phys. 87, 345 (1992).
[3] J.M. Kosterlitz and D J . Thouless, J. Phys. C 6, 1181 (1973).
[4] D.A. Bell, H.F. Hess, G.P. Kochanski, S. Buchman, L. Pollack,
Y.M. Xiao, D. Kleppner, and TJ. Greytak, Phys. Rev. B 34, 7670
(1986).
[5] L. Pollack, S. Buchman, and T.J. Greytak, Phys. Rev. B 45, 2993
(1992).
[6] JJ. Berkhout, E.J. Wolters, R. van Roijen, and J.T.M. Walraven,
Phys. Rev. Lett. 57, 2387 (1986).
[7] M.W. Reynolds, I.D. Setija, and G.V. Shlyapnikov, Phys. Rev. B 46,
575 (1992).
[8] S.A. Vasilyev, E. Tjukanov, M. Mertig, A. Ya. Katunin, and
S. Jaakkola, Europhys. Lett. 24, 223 (1993).
[9] J. Helffrich, M.P. Maley, M. Krusius, and J.C. Wheatley, Phys. Rev.
B 34, 6550 (1986).
[10] D.O. Edwards, I.B. Mantz, and V.U. Nayak, J. Phys. (Paris), Colloq.
39, C6-300 (1978).
[II] I.B. Mantz, D.O. Edwards, and V.U. Nayak, Phys. Rev. Lett. 44,
663 (1980).
[12] Yu. Kagan, B.V Svistunov, and G.V. Shlyapnikov, Sov. Phys. JETP,
66, 314 (1988).
[13] H.T.C. Stoof and M. Bijlsma, preprint (1993).
23
BEC of Biexcitons in CuCl
Masahiro Hasuo and Nobukata Nagasawa
Department of Physics
The University of Tokyo
7-3-1 Hongo, Bunkyo-ku, Tokyo, 113
Japan
Abstract
A new approach to detecting coherence associated with Bose-Einstein con-
densation of biexcitons at K=0 in CuCl is described. Remarkable enhance-
ment of the phase-conjugate signal at the two-photon resonance of the biex-
citons at K=0 is observed when high-density incoherent biexcitons are in-
jected, suggesting macroscopic increase of the occupation number of the
K=0 biexciton state and enhancement of the coherence. A small blue shift
of the K=0 biexciton resonance is also observed, suggesting the change
of the chemical potential of the biexciton due to the interaction between
biexcitons.
1 Introduction
An excitonic particle such as an exciton or a biexciton (excitonic molecule)
in a semiconductor crystal has a boson-like nature. If the system of
excitonic particles is regarded as an ideal Bose gas, Bose-Einstein con-
densation (BEC) may occur at the critical density of
_ 2.612 2mkBT 3/2
c { ] U)
~ J4^ ~W~ '
where m and T are the effective mass of the particle and the temperature
of the system, respectively.
In most semiconductors, the biexciton, a bound complex of two exci-
tons, is the most stable entity, with the lowest energy per electron-hole
pair. The CuCl crystal is famous as such a material, the excitonic param-
eters of which have been well determined experimentally. For example,
the effective mass of the biexciton is 5.3 times as large as the electron rest
487
488 M. Hasuo and N. Nagasawa
mass [1]. Thus, the corresponding Nc is estimated to be 2.2 x 1017/cm3
at T=2 K.
A biexciton in CuCl decomposes into one photon and one exciton
through a radiative process that gives characteristic luminescence called
M-emission. The spectral shape of the M-emission is well known to be
determined by the density of states and the distribution function of the
biexciton system. When the density of biexcitons, N, is far less than
Nc, the spectral shape is reproduced approximately by the Maxwell-
Boltzmann distribution function [2]. With increase of JY, it is expected
that the spectral shape changes, reflecting the Bose statistics due to the
Bose nature of the biexciton.
In fact, it has been reported that the spectral shape of M-emission
is reproduced by the Bose-Einstein distribution function corresponding
to fx=O under special experimental conditions, where \i is the chemical
potential of biexcitons [3]. A demonstration showing the quantum at-
traction between biexcitons in phase space has also been presented by
the use of a pump-and-probe method on the M-emission associated with
the two-photon excitation of biexcitons [4].
When the BEC occurs, it is expected that a macroscopic number of
biexcitons collapses to a single quantum state of zero momentum, K=0,
and forms a coherent quantum state. However, the coherence inherent in
the BEC has never been demonstrated experimentally so far. Recently,
we have proposed a new pump-and-probe method that aims to detect
the coherence [5].
We measure the change in the intensity of the phase-conjugate signal
generated from the third-order macroscopic non-linear polarization asso-
ciated with the coherent biexcitons populated at K=0 under the presence
of the high-density biexcitons injected incoherently. Fig. 1 shows a
schematic illustration of the phase conjugation. The relevant non-linear
polarization is expressed by
P^^^(Ei'E2)Em39 (2)
where E\ and Ej are the electric fields of the counter-propagating laser
beams 1 and 2 used for exciting the coherent biexcitons at X=0 [6].
Adjusting these beams' intensities, the density of the coherent biexcitons
populated at K=0 is substantially lower than Nc. The electric field £3 of
the laser beam 3 is used to induce a parametric scattering process associ-
ated with the coherent biexcitons, as shown in Fig. 1. The signal light field
has phase which is conjugate to that of £3, as seen in Eq. (2). Since this
phase-conjugate signal is generated from the coherent assembly of the
BEC ofBiexcitons in CuCl 489
(a)
Beam! Beam2
Beam3
' Signal
(b)
- lbiexciton>
Beam2 Beam3
EGTA EGTA + 6
Beami Signal
EGTA EGTA - 6
Fig. 1. Geometry (a) and energy diagram (b) for the phase conjugation.
relevant microscopic polarizations which enter into %(3), one can monitor
the occupation number of the biexcitons at K=0 and its coherence. If
BEC is realized, it is expected that the intensity of the phase-conjugate
signal should show a remarkable increase as an indication of the onset
of BEC and of the enhancement of the relevant coherence.
2 Experimental Method
Figure 2 shows a schematic diagram of the experimental set-up. Each
photon energy of the counterpropagating beams 1 and 2, provided by
Dye Laser A, was adjusted to be the two-photon resonance energy of the
biexcitons at K=0, £Gr^=3.1858 eV. The spectral width of the laser light,
7i,2> was adjusted to be 138 jueV for a reason mentioned below. These
beams were focused in a sample. The spot size on the sample surface was
about 100 /xm</>, resulting in a total intensity of 160 kW/cm2. The phase
490 M. Hasuo and N. Nagasawa
Beam 1
ND--
Monochromator • «— ^
P
HM /
ND Q
Dye 1 aser B
Beam 3 \/
Dye 1 aser C
Beam 4
Fig. 2. Schematic diagram of the experimental set-up. PBS :polarized beam
splitter, ND:neutral density filter, Qiquarter wave plate, P:polarizer, VA:variable
attenuator, HM :half mirror. Inset shows a phase-conjugate spectrum expected
under the present excitation conditions. For details, see text.
where AK and L are the relevant phase mismatch [7] and the sample
thickness, respectively. The dashed curve shows the phase-matching fac-
BEC of Biexcitons in CuCl 491
tor as a function of photon energy estimated for the present experimental
conditions.
The polarization of these laser beams is made circular so as to excite
mainly the K=0 biexcitons via the relevant selection rule of the two-
photon transition. The external angle between beams 1 and 3 was 10°,
which corresponds to an internal angle of 3.6°. The phase-conjugate
signal was analyzed with a high resolution monochromator (Jobin-Yvon
THR1500, band pass 16/ieV).
The fourth laser beam, from Dye Laser C, was used to generate the
incoherent biexcitons through a bimolecular collision process of photo-
excited excitons. We shall refer to this light as the incoherent pump
light hereafter. The photon energy and the spectral width of this light
were 3.2029 eV and 1.36 meV, respectively. The beam was focused on
the sample surface to a spot of 200 fimcj) in order to cover the region
illuminated by beams 1, 2 and 3. The intensity of this light was adjusted
continuously with a variable intensity attenuator. The maximum intensity
was 2.7 MW/cm2.
All the dye lasers (BBQ in p-dioxane) were pumped by a common
excimer laser (Lambda Physik 53MSC). The pulse duration of each dye
laser light was about 10 ns.
The sample was a 15 /mi thick single-crystal platelet, grown from vapor
phase. All the measurements were done at 2 K.
"CO
a>
0 -
dN
n 2 (4)
ym
~dt 2
dn 9
_ = -rjn1 + ymo\N - yexn + J, (5)
at
BEC ofBiexcitons in CuCl 493
0.0
500 1000 1500 2000 2500 3000
2
Intensity (kW/cm )
J = (6)
LeSE
Here, R,I and E are the reflectivity, intensity and photon energy of the
incoherent pump light, respectively; Leff is the effective excitation depth
and was assumed to be expressed in the form
Leff = (7)
where vg and ysc are the group velocity and the scattering rate of the
exciton polaritons which are converted from the incoherent pump light
at the sample surface. They were adopted to be 1.5 x 10 6cm/s and
1 x 109s~1, respectively [10]. The collision rate of excitons to create the
incoherent biexcitons is rjn. As seen in Fig. 4, the remarkable increase
494 M. Hasuo and N. Nagasawa
102
Biexciton density ( c m 3 )
4 Conclusion
We have presented a new approach for observing the coherence associ-
ated with the BEC of the biexciton system in CuCl. A phase-conjugate
signal was used as a probe to monitor the macroscopic occupation of
the biexcitons at K=0 and the relevant coherence under the high-density
injection of the biexcitons of incoherent nature. A remarkable increase of
BEC ofBiexcitons in CuCl 495
the signal intensity at the critical density for BEC and a blue shift of the
K = 0 biexciton resonance were observed. These results are attributed to
the presence of the Bose-Einstein condensate of biexcitons.
References
[I] T. Mita, K. Sotome and M. Ueta, J. Phys. Soc. Jpn. 48, 496 (1980).
[2] H. Haug, and E. Hanamura, Physics Reports 33, 209 (1977).
[3] L.L. Chase, N. Peyghambarian, G. Grynberg, and A. Mysyrowicz,
Phys. Rev. Lett. 42, 1231 (1974).
[4] N. Peyghambarian, L.L. Chase, and A. Mysyrowicz, Optics Comm.
41, 178 (1982).
[5] M. Hasuo, N. Nagasawa, and A. Mysyrowicz, Phys. Stat. Sol.(b) 173,
255 (1992); M. Hasuo, N. Nagasawa, T. Itoh, and A. Mysyrowicz,
Phys. Rev. Lett. 70, 1303 (1993).
[6] G. Mizutani, and N. Nagasawa, J. Phys. Soc. Jpn. 52, 2251 (1983);
L.L. Chase, M.L. Claude, D. Hulin, and A. Mysyrowicz, Phys. Rev.
A 28, 3969 (1983).
[7] Y. Masumoto, and S. Shionoya, J. Phys. Soc. Jpn. 49, 2236 (1980).
[8] T. Hatano, doctoral thesis (The University of Tokyo, 1993).
[9] H. Akiyama, T. Kuga, M. Matsuoka, and M. Kuwata-Gonokami,
Phys. Rev. B 42, 5621 (1990).
[10] T Ikehara, and T. Itoh, Phys. Rev. B 44, 9283 (1991).
[II] L.V. Keldysh, and A.N. Kozlov, Sov. Phys. JETP 27, 521 (1968).
[12] M. Inoue, and E. Hanamura, J. Phys. Soc. Jpn. 41, 1273 (1976).
24
The Influence of Polariton Effects on BEC
of Biexcitons
A. L. Ivanov and H. Haug
Institut fur Theoretische Physik
J.W. Goethe-Universitat
Robert-Mayer-Str. 8, D-60054 Frankfurt/Main
Germany
Abstract
The influence of the radiative decay of excitonic molecules on a possible
quasiequilibrium Bose-Einstein condensation (BEC) of excitonic molecule's
is examined with respect to the radiative renormalization of the excitonic
molecule energy (excitonic molecule Lamb shift). For the excitonic molecule
wave function, a Schrodinger equation which contains polariton effects is
derived and analyzed. Both the inverse excitonic molecule radiative lifetime
ym and the biexciton Lamb shift Am depend strongly on the total excitonic
molecule momentum K. The energy renormalization Am(K) leads to the ex-
citonic molecule effective mass modification and can result in a camel-back
structure at K = 0, which opposes a BEC of excitonic molecules at K= 0.
1 Introduction
Observations of a quasiequilibrium Bose-Einstein condensation (BEC)
of excitonic molecules (EM) have been attempted [1, 2, 3] following
its theoretical prediction [4, 5] (for reviews see, e.g. [6, 7]). Recent
approaches [8, 9] with high-precision techniques renewed the interest in
this phenomenon. Traditionally, one tries to detect BEC of the EMs
in luminescence. In the new approach the appearance of coherence in
the thermal system of the EMs has been tested by means of four-wave
mixing and treated as a fundamental manifestation of BEC. (See the
paper by Hasuo et al. in this book.) Both optical methods for the
BEC detection imply that the optical transition to the corresponding
intermediate exciton (IE) state is dipole-active. In this case the EM state
is unstable against optical decay with a "giant" oscillator strength [6].
The short radiative lifetime rm of the EM is a very serious obstacle for
496
The Influence of Polariton Effects on BEC of Biexcitons 497
BEC. The influence of this factor has been considered exclusively in the
kinetics of the Bose condensate formation from arbitrary initial non-
equilibrium distributions of EMs (see, e.g. [10]). The strong interaction
with electromagnetic field also modifies the dispersion of the elementary
excitations, however. A well-known example is the Lamb shift of the
resonance energy of a dipole-active two-level system. Another example
is the polariton spectrum of dipole-active IEs. Here we study the EM
Lamb shift and show that this effect strongly influences the BEC of EMs.
Here, p and K are the momenta of the relative and center-of-mass motion
of the EM, respectively. H*(k) = k2/2M is the kinetic energy of an IE,
and W\2(p — Pi) is the attractive potential between two singlet IEs with
opposite internal electron spin orientation. Due to the quadratic IE
dispersion, the center-of-mass motion splits off, and the wave function
and energies of the relative motion are independent of K. This equation
in the underlying e-h picture has often been used to calculate the EM
properties variationally [11, 12, 13, 14]. In order to investigate the
influence of the polariton effects on the interior structure of EM we
derived self-consistently [15] an EM wave equation from the complete
e-/i-photon (e — h — y) Hamiltonian :
£ [{HP (p + f) + H<> (-p + f)) <5p,Pl + Wn (p, Pl ,K)]
Pi
with Hp(k) = co~(k), where co±(k) is the dispersion of the upper (+) or
m-
lower (-) polariton branch, which are given by the roots of
2
n (3)
€Q(O2 CO2 + COtp2/M — CO2
Here, coy(p) is the photon frequency and eo is the background optical
dielectric constant. The polariton parameter Qc is defined in terms of
498 A. L. Ivanov and H. Haug
the longitudinal-transverse (t — t) splitting co/f and the transverse exciton
frequency cot:
Q2C = 2cott(ot. (4)
In Eq. (2) only the lower polariton dispersion branch is taken into
account. In the following we will consider only the ground states of the
IE and EM (J=l). The linear equation, Eq. (2), contains the radiative
decay of an EM state, as well as the renormalization of the EM energy
(EM Lamb shift) due to the polariton effects. When an IE in the
EM acquires a small momentum within the optical range (sectors A1A2
and B1B2 in Fig. 1), the EM undergoes a radiative annihilation. In
other words, the linear part of the lower polariton dispersion branch
corresponds to the case without a bound EM state. The radiative decay
of an EM stems from the polariton dispersion of its IE rather than from
the usual tunneling process out of a metastable state. In our picture
the EM luminescence is a continuous evolution of the EM interior state
rather than a discrete act of the EM optical conversion to y and IE.
3 Biexciton Lamb-Shift
The non-stationary solution of Eq. (2), ^(p), represents an outgoing
spherical wave due to the radiative decay of a EM state with Q£ =
Q£ + Am(K) - iym(K). Here, Q£ = Q£ =o + K2/2Mm is the unperturbed
EM energy with the EM translational mass Mm = 2M. For small relative
The Influence of Polariton Effects on BEC of Biexcitons 499
03
GO
momenta which belong to the optical range of the polariton spectrum, the
EM wavefunction (EMWF) ^(p) is strongly modified. This modification
of the EMWF cannot be treated perturbationally. Here the EM Lamb
shift Am(K) and the inverse radiative lifetime ym(K) are of the same order,
while in atomic optics the radiative energy shifts are always substantially
smaller than corresponding inverse lifetimes. In order to treat Eq. (2) we
decompose the EMWF in
where *F(p) is the known solution of unperturbed Eq. (1), while *Fi(p,K)
describes the influence of the polariton effects and is large only in the
optical range, that is, for p « a~\ where a m is the EM radius. For these
small momenta one neglects the momentum dependence of the attractive
potential Wn{\> — Pi) = Wnity =const.
Substituting Eq. (9) in Eq. (2), one finds the following integral equation
for ^ ( p ) in a D-dimensional system:
(13)
(14)
The functions *F(p) and ^(pjK), as well as the basic Eq. (2), are symmet-
ric with respect to the substitution p —• —p. According to Eq. (12), the
roots po = Po(K) of the equation A(p, K) = 0 give rise to singularities in
^ ( P J K ) . These singularities describe the two outgoing polariton waves
with momenta po 4- K/2 and — po + K/2 in the decay of the EM with
K. A(p,K) = 0 is the energy-conservation law of this decay. The graphic
solution of this equation is presented in Fig. 2.
According to Eq. (2) the total EM momentum K influences the relative
motion of the IEs in the EM. The polariton dispersion makes *Fi(p,K),
Am(K), and ym(K) K-dependent in the optical range K < 2po(K = 0).
Formally, the solution (12) can be found for an arbitrary EM energy
Q£. In order to find the true EM energy for a given K, we change
Q£ continuously around the unperturbed EM energy Q£ and find the
value which minimizes the outgoing part / \x¥i(p,K)\2dDp/(2n)D and
maximizes the EM radiative lifetime. Although the wave functions (12)
cannot be normalized, one can compare the relative contributions of the
outgoing parts. For simplicity we present here only numerical solutions
for one-dimensional EM systems with D = 1. The fraction of the phase
space in which polariton effects dominate over the Coulomb IE-IE
interaction is of the order of (ampo(K = 0))D. Therefore, the influence
of polariton effects on the EMWF is much stronger in one dimension
than in three dimensions. We use the parameters of GaAs/GaAlAs
quantum-well wires [16]. For the attractive IE-IE potential
The Influence of Polariton Effects on BEC of Biexcitons 501
m
Fig. 2. Graphic solution of Eq. (12). The points 0 and O\ correspond to outgoing
polaritons.
we use the simple form of the deuteron model [17] in dimensionless units
(normalized with the EM Rydberg and the EM radius)
Wn(p-Pi) = (15)
For P = 4.0, the parameter a = 0.5 minimizes the energy and gives
the binding energy em — 0.75. In dimensional units these parameters
correspond to em = -5 meV and EM Bohr radius am = 73.9A. The
accuracy of the variational procedure is about 2.5% [17].
502 A. L. Ivanov and H. Haug
100
50
-1.0
On r3
CD
-10-
-15
30
K (xlO cm"1)
Fig. 4. Biexciton energy AO£ = X2/2Mm + Am(K) (full line) and inverse radiative
lifetime ym(K) (dashed line) versus total momentum K. All parameters are the
same as in Fig. 3, except for yx = 500 cm"1.
4 Discussion
We expect that in a three-dimensional system such as bulk CuCl a
camel-back or, rather, "Mexican hat" structure will also be obtained
for the renormalized energy, particularly because the heavy hole mass
of this material results in a heavy EM mass of Mm = 5.2mo, which
favors the minimum at Ko. The magnitude of the minimum Am(X0),
however, is expected to be considerably smaller than in one dimension.
The appearance of this minimum will eliminate the possibility of a BEC
of EMs at K = 0. But at the same time a condensation in the minimum
at KQ ~ 2po(K = 0) (see Fig. 5) is not possible, because the relation for
The Influence of Polariton Effects on BEC of Biexcitons 505
the total number of EMs
yields for all finite temperatures /x(iVm, T) < 0, if we scale the EM min-
imum energy at Ko to be zero. Our results are consistent with the
experimental data of Refs. [2, 3]. In these experiments a strong accumu-
lation of the EMs at K ~ 2po(K = 0), rather than at K = 0, has been
detected.
References
[I] N. Nagasawa, T. Mita, and M. Ueta, J. Phys. Soc. Jpn. 41, 929
(1976).
[2] L.L. Chase, N. Peyghambarian, G. Gryndberg, and A. Mysyrowicz,
Phys. Rev. Lett. 42, 1231 (1979).
[3] N. Peyghambarian, L.L. Chase, and A. Mysyrowicz, Phys. Rev. B
27, 2325 (1983).
[4] E. Hanamura and M. Inoue, in Proc. 11th International Conference
on the Physics of Semiconductors, Warsaw, (1972) p. 711.
[5] S.A. Moskalenko, Fiz. Tverd. Tela 4, 276 (1962)
[6] E. Hanamura and H. Haug, Physics Reports 33, 209 (1977).
[7] A. Mysyrowicz, J. de Physique (Supplement) 41, C -281 (1980).
[8] M. Hasuo, N. Nagasawa, and A. Mysyrowicz, Phys. Stat. Sol. (b)
173, 255 (1992).
[9] M. Hasuo, N. Nagasawa, T. Itoh, and A. Mysyrowicz, Phys. Rev.
Lett. 70, 1303 (1993).
[10] M. Ueta, H. Kanzaki, K. Kobayashi, Y. Toyozawa, and E. Hana-
mura, Excitonic Processes in Solids (Springer-Verlag, Berlin, 1986),
Springer Series in Solid-State Sciences 60, p. 110.
[II] O. Akimoto and E. Hanamura, Solid State Comm. 10, 253 (1972).
[12] A.I. Bobrysheva, M.F. Miglei, and M.I. Shmiglyuk, Phys. Stat. Sol.
(b) 53, 71 (1972).
[13] W.F. Brinkman, T.M. Rice, and B. Bell, Phys. Rev. B 8, 1570 (1973).
[14] M.I. Sheboul and W. Ekardt, Phys. Stat. Sol. (b) 73, 165 (1976).
506 A. L. Ivanov and H. Haug
[15] A.L. Ivanov and H. Haug, Phys. Rev. B 48, 1490 (1993).
[16] L. Banyai, I. Galbraith, C. Ell, and H. Haug, Phys. Rev. B 36, 6099
(1987).
[17] S. Fliigge, Practical Quantum Mechanics (Springer, Berlin 1974), p.
196.
[18] H. Akiyama, T. Kuga, M. Matsuoka, and M. Kuwata-Gonokami,
Phys. Rev. B 42, 5621 (1990).
[19] M. Hasuo, N. Nagasawa, and T. Itoh, Opt. Commun. 85, 219 (1991).
[20] T. Ikehara and T. Itoh, Sol. Stat. Comm. 79, 755 (1991).
25
Light-Induced BEC of Excitons and
Biexcitons
A. I. Bobrysheva and S. A. Moskalenko
Institute of Applied Physics
Academy of Sciences of Moldova
5 Academy Street, Kishinev
Moldova
Abstract
We review the theory of coherent pairing of excitons and biexcitons in two-
dimensional and three-dimensional semiconductor structures.
1 Introduction
The coherent pairing of bosons, formally analogous to Cooper pairing
of electrons in superconductors, and the coexistence of particle and pair
Bose-Einstein condensation (BEC) was first studied in Refs. [1-4]. BEC
of excitons in semiconductors was investigated using the approximation
of coherent pairing of the electrons and holes [5]. Later on, the idea of
coherent pairing of bosons was applied to excitons [6-8]. In Ref. [6],
it was assumed that the boson pairs consist of two electron-hole pairs
instead of two excitons, and it was suggested that the new collective
state is a BEC even though the two-pair bound state does not exist. It
was proved independently [7, 8] that in a system of bosons and excitons
with an attractive pair interaction sufficient for biexciton formation, the
coherent pairing of excitons with momenta k and —k coincides with the
BEC of biexcitons with zero translational momentum.
The possibility of particle and pair BEC in a system of four species
of excitons with different values of spin projections of the electron and
hole was studied in Ref. [9]. It is known [10, 11] that the interaction of
an exciton of either type with other exciton species is repulsive on the
average, but that pairs of excitons with antiparallel spins of electrons and
holes interact attractively, and biexcitons can be formed. In two limiting
cases, when the exciton-exciton interaction energy W is much greater or
smaller than ortho-paraexciton splitting A0_p, it was shown that the BEC
507
508 A. I. Bobrysheva and S. A. Moskalenko
of biexcitons persists without collapse due to the internal spin structure
of excitons. In the case where A0_p < W the biexciton is formed by
all four species of excitons. When A0_p > W in a system of para-
and orthoexcitons, BEC of paraexcitons and biexcitons formed by two
orthoexcitons with zero total spin can take place simultaneously. Taking
the CuCl crystal as an example, it was proved [12] that in the presence of
nonresonant polarized laser radiation, which induces the polarized exci-
tonic BEC, coherent pairing of excitons takes place. Two induced BECs
coexist: that of polarized excitons with wave vector ko, and the biexci-
tonic one with wave vector 2ko. It was shown that the biexciton becomes
anisotropic in the sense that four species of excitons T$x, Tsy, r$z and
F2 do not participate equally in biexciton formation, but have different
weights. That is, the two-photon biexciton absorption becomes polarized.
In the present paper, this phenomenon is studied in a wider class of
crystals with different dimensionalities d = 2,3, in particular in quasi-
two-dimensional semiconductor structures with quantum wells such as
GaAs/AlxGai_xAs and ZnSe/(Zn,Mn)Se and three-dimensional semi-
conductor crystals such as CuCl. Optical methods which allow the
observation of the new electronic state of the crystal are of interest.
Therefore biexciton hyper-Raman scattering and luminescence under the
above-mentioned conditions are also studied.
Hex_ex = 4
ki,k2,q tr
(2)
510 A. I. Bobrysheva and S. A. Moskalenko
Hex-ex has the form given by (2). We assume that laser radiation is
polarized along the x axis and excites the exciton mode E(x) with
the wave vector ko; Hex-R describes this exciton-electromagnetic field
interaction.
The induced BEC of polarized excitons in the state ko is introduced
in the Hamiltonian by Bogoliubov's translation operation. The biexciton
creation operator has the form
where the function of relative motion O(q) and the coefficients Q obey
normalization conditions. Therefore, BEC of biexcitons in the state 2ko is
introduced by the operation of coherent pairing of excitons (ko+q,ko—q)
of the same type i.
The coefficients of the (uuvt) transformations are given by
= (|< v >\ TA)(y/2^3(\< v >\ +A)2 + (|< v >\ -A) 2 )" 1 , (8)
Light-Induced BEC of Excitons and Biexcitons 511
where 1, 2 refer to signs minus, plus, respectively; < v > is the mean
potential energy of the internal motion; A is the biexciton detuning. It
follows from Eq. (7) that for A ^ 0, the number of excitons of each of the
four types forming the biexciton are different. In this sense the biexciton
is polarized. At A = 0, fast mutual conversions of the excitons remove
this inequality.
Abstract
513
514 I. V. Beloussov and Yu. M. Shvera
The investigation reported in [5] was based on a model formally analo-
gous to that used by N.N. Bogoliubov in Ref. [6] to study the equilibrium
system of a weakly non-ideal Bose gas. In the non-equilibrium situation
considered in [5], in which decay of the polariton condensate and exci-
tation of non-condensate polaritons take place, this model is adequate
for the experimental situation only at the initial stage of the condensate
decay, when the number of condensate polaritons is still much greater
than the total number of non-condensate polaritons. The system is es-
sentially non-stationary, while a real energy spectrum implies a steady
state of the system (see e.g. Ref. [7], p.46). Thus the results of Refs. [4, 5]
for the polariton energy spectrum based on the above-mentioned model
are only approximate.
Because of the essential non-stationarity of the processes in the system,
the methods of non-equilibrium statistical mechanics must be used to
describe it adequately. The derivation of equations that describe the
kinetics of the polariton condensate decay and the excitation of quantum
fluctuations have some specific features due to the degeneracy in the
system. As the energy and wavevector of two non-condensate polaritons
can be equal to the energy and wavevector of two condensate polaritons,
respectively, there is degeneracy of two-particle states. Moreover, the
presence of the condensate in the system also leads to degeneracy due to
its macroscopic amplitude [8].
The correct description of the system with degeneracy requires the
introduction of abnormal (or off-diagonal) distribution functions [9]
an<
l ^k = (^k^ko-k), together with the usual (normal) ones
). Here ®jj~(<l>k) are Bose operators of creation (annihilation)
of a polariton on the lower branch with wave vector k. The appearance of
abnormal averages and the coherent part of the polariton field indicates
a breaking of the selection rules connected with the gauge invariance
of the system [9-11]. It can take place as a result of the action of
external classical sources, spontaneously, or because of non-invariant
initial conditions due to the action of external sources.
An earlier attempt to obtain kinetic equations for polaritons excited in
semiconductors by an external classical field has been given in Refs. [12,
13]. However, in this work the degeneracy of two-particle states was
not taken into account, and the abnormal distribution functions Ft were
not introduced. As a result, one expects that the equations obtained in
Refs. [12, 13] will lead to unphysical singularities.
Kinetic equations describing the evolution of partially coherent polari-
tons which take into account the degeneracy were obtained in [14, 15]
Evolution of a Nonequilibrium Polariton Condensate 515
using the Keldysh method [16], presented in terms of functionals. They
coincide with the equations obtained in [17] using the non-equilibrium
statistical operator method [J8] and do not possess unphysical singu-
larities. One can find extended versions of Refs. [14, 15, 17] in Refs.
[19-21].
According to Refs. [15, 17], the kinetics of partially coherent polaritons
are described in the Born approximation by a closed set of nonlinear
integro-differential equations for the coherent part of polariton field ^ ^
and the normal n^ = N^ — ^kj^kol 2 and abnormal /k = F^ — ^Fj^
distribution functions. In the absence of quantum fluctuations described
by the functions n^ and /k, the equations become identities, and the
equation for ^Fka (a = 1,2 is the label of the polariton branch) coincides
with that obtained in Ref. [22] for a interacting system of coherent
excitons and photons. In the case when *Fko = 0 and /k = 0, the
equations obtained in [15, 17] reduce to the usual kinetic equation for
the distribution function N^ (see e.g. Ref. [23]).
The right-hand sides of the equations obtained in [15, 17] include terms
linear in the exciton-exciton interaction strength v, and terms propor-
tional to v2. Terms proportional to v correspond to the self-consistent
field approximation (SCFA), which neglects the higher correlation func-
tions. This approximation is sufficient to describe the initial stage of
evolution. It is shown [15] that depletion of the coherent part ^F^ a n d
excitation of quantum fluctuations n^ and / ^ in the condensate mode
occurs at this stage. Thus it partially loses its coherence. There exist two
regions of wave vector k ^ ko. In the first one, describing the instability
region, the number of polaritons increases monotonically; in the second
one, it is a periodic function of time with an amplitude decreasing with
the distance from the instability region.
As usual, the terms proportional to v2 describe the difference between
the number of processes (per unit of time) of polariton creation and
annihilation in a state with wavevector k. They only arise if non-
condensate polaritons exist in the system. They describe the evolution
which is slower than that described by terms proportional to v, which
correspond to resonant scattering with participation of two condensate
polaritons.
It should be mentioned that the SCFA takes into account the influence
of the excited non-condensate polaritons on the condensate. On the
other hand, it describes the fastest processes in the system. So one should
expect the establishment of steady state in the isolated polariton system
after the time interval T ~ h/vn (n is the condensate initial density).
516 I. V. Beloussov and Yu. M. Shvera
Terms in the kinetic equations proportional to v2 are responsible for its
further evolution.
Using considerations similar to those used in [24], we have found the
exact steady-state solution of the evolution equations given by the SCFA.
Formally, it has the following nature. In the Heisenberg representation,
the SCFA corresponds to the Hamiltonian [15] which has operator
terms O£^io-k w ^^ time-dependent coefficients £ = ^F^ + A (here
A = J2kfa)- A t t h e initial stage of evolution, when I^FkJ2 > |A|,
these terms are responsible for the appearance and development of the
polariton condensate instability [4, 5]. In the steady state, the two terms
in £, cancel each other, so that { = 0. Thus these terms disappear from
the Hamiltonian.
We note that in [4, 5], the influence of non-condensate polaritons on the
condensate was not taken into account, and the term A in the expression
for £ did not appear. As a result, in [4, 5] the possibility of obtaining a
steady-state solution, which is the result of mutual compensation of ^Fj^
and A, did not exist.
From the physical point of view, the results obtained here mean
that while the condensate is being depleted, and polaritons are being
excited in the instability region (modified by the concentration-dependent
corrections), the backward processes (in which scattering of two non-
condensate polaritons creates two polaritons in the condensate) become
more important. It results in slowing down of the parametric instability,
and finally in the establishment of a steady state. This state corresponds
to a dynamic equilibrium between the condensate and non-condensate
polaritons.
The steady state is characterized by the renormalized frequency of the
polariton condensate hco^ = hQ,^ + yjlvno and renormalized energies of
non-condensate polaritons E^ = hQ^ + 2vno (here Q^ is the frequency
of non-interacting polaritons). This result differs crucially from that
obtained in [4, 5]. Distribution functions of non-condensate polaritons
are localized in the regions of k-space, as determined by the resonant
condition E^ + £2ko-k — 2fta>ko = 0-
Future investigations of this problem require numerical analysis of the
set of integro-differential equations corresponding to the SCFA. Another
important problem, in our opinion, is the study of polariton kinetics
under the action of an external field.
Evolution of a Nonequilibrium Polariton Condensate 517
We wish to thank Dr. A. Ivanov and Dr. S. Tikhodeev for helpful
discussions.
References
[I] A.L. Ivanov and L.V. Keldysh, Zh. Eksp. Teor. Fiz. 84, 404 (1983).
[2] M.V. Lebedev, Zh. Eksp. Teor. Fiz. 101, 957 (1992).
[3] S.A. Moskalenko, Introduction to the Theory of High Density Exci-
tons (§tiinja, Kishinev, 1983).
[4] M.I. Shmigliuk and V.N. Pitei, Coherent Polaritons in Semiconductors
(§tiin{a, Kishinev, 1989).
[5] V.N. Pitei, M.I. Shmigliuk, V.T. Zyukov, and M.F. Miglei, in Bose
Condensation of Polaritons in Semiconductors, pp. 58-79 (§tiinja,
Kishinev, 1985).
[6] N.N. Bogoliubov, J. Phys. USSR 11, 23 (1947); see also N.N. Bogoli-
ubov, Selected Works. Vol.2, pp. 242-257 (Naukova Dumka, Kiev,
1971).
[7] L.D. Landau and E.M. Lifshitz, Quantum Mechanics (Pergamon,
Oxford, 1963).
[8] J.R. Schrieffer, Theory of Superconductivity (Nauka, Moscow, 1970).
[9] N.N. Bogoliubov, Quasi-Averages in the Problems of Statistical
Mechanics. Dubna: Preprint of JINR R-1415, 1963; see also N.N.
Bogoliubov, Lectures on Quantum Statistics (Gordon and Breach,
NY, 1970), Vol. 2.
[10] N.N. Bogoliubov (Jr) and B.I. Sadovnikov, Some Questions of Sta-
tistical Mechanics (Vysshaya Shkola, Moscow, 1975).
[II] V.N. Popov and VS. Yarunin, Collective Effects in Quantum Statistics
of Radiation and Matter (Izd. LGU, Leningrad, 1985).
[12] S.A. Moskalenko, A.H. Rotaru and Yu.M. Shvera, in Laser Optics of
Condensed Matter, J.L. Birman, H.Z. Cummins and A.A. Kaplyan-
skii, eds. (Plenum, NY, 1988), pp. 331-336.
[13] VR. Misko, S.A. Moskalenko, A.H. Rotaru and Yu.M. Shvera,
Phys. Stat. Sol. (b) 159, 477 (1990).
[14] I.V Beloussov and Yu.M. Shvera, Phys. Stat. Sol. (b) 139, 91 (1990).
[15] I.V Beloussov and Yu.M. Shvera, Z. Phys. B 90, 51 (1993).
[16] L.V. Keldysh, Zh. Eksp. Teor. Fiz. 47, 1515 (1964).
[17] I.V. Beloussov and Yu.M. Shvera, Teor. Mat. Fiz. 85, 237 (1990).
[18] D.N. Zubarev, Non-Equilibrium Statistical Thermodynamics (Plenum,
NY, 1974).
518 I. V. Beloussov and Yu. M. Shvera
[19] I.V. Beloussov and Yu.M. Shvera, in Excitons and Biexcitons in
Confined Systems (§tiinja, Kishinev, 1990), pp.133-176.
[20] I.V. Beloussov and Yu.M. Shvera, in Interaction of Excitons with
Laser Radiation (§tiinja, Kishinev, 1991), pp.57-74.
[21] I.V. Beloussov and Yu.M. Shvera, in Nonlinear Optical Properties
of Excitons in Semiconductors of Different Dimensionalities (§tiinja,
Kishinev, 1991), pp.53-64.
[22] L.V. Keldysh, in Problems of Theoretical Physics (Nauka, Moscow,
1972), pp.433-444.
[23] A.I. Akhiezer and S.V. Peletminskii, Methods of Statistical Physics
(Nauka, Moscow, 1977).
[24] V.L. Safonov, Physica A 188, 675 (1992).
27
Excitonic Superfluidity in CU2O
E. Fortin and E. Benson
Department of Physics
University of Ottawa
Ottawa, ON KIN 6N5
Canada
A. Mysyrowicz
LOA, Ecole Poly technique
Palaiseau
France
Abstract
519
520 E. Fortin, E. Benson and A. Mysyrowicz
Au electrode
Photocurrent
^ ^ 0 . 2 5 4 mm
Cu electrode
1 mm
f1 1 • 1\1
E|
|
Exciton
formation E>
Cuprite crystal
>
Laser
beam
Fig. 1. Schematic of the experiment. The sample is oriented with the < 100 >
direction along the edges.
Thi( dcness = 3 56 mm
I
1 V
1-
1
1
Jft m f /
0.00
Time (psec)
J
rj^r\Kj\/\r^KJ^\r^\^^ ^^
We believe that these results are strong evidence for the onset of a
superfluid excitonic phase above a critical density and below a critical
temperature. Let us first compare the density of excitons inside the
initial packet to the critical density for Bose-Einstein condensation as
predicted by the ideal Bose model. Since the measurements are taken
522 E. Fortin, E. Benson and A. Mysyrowicz
Tl = 2 2 ' 3 = 1.58
T
This is in good agreement with the experimentally observed value of 1.55.
In conclusion, detection of exciton transport with good time and
space resolution reveals the onset of an anomalous ballistic transport
below a critical temperature and above a critical particle density. The
conditions required for the appearance of ballistic transport are in very
good agreement with those predicted for Bose-Einstein condensation in
an ideal Bose gas. We interpret these results as evidence for a superfluid
excitonic phase associated with Bose-Einstein condensation. For more
details, see Ref. [6].
References
[1] E. Tselepis, E. Fortin and A. Mysyrowicz, Phys. Rev. Lett. 59, 2107
(1987).
[2] A. Mysyrowicz, D. Hulin and A. Antonetti, Phys. Rev. Lett. 43,
1123 (1979); D. W. Snoke, A. J. Shields and M. Cardona, Phys. Rev.
B45, 11693 (1992).
[3] A. Mysyrowicz and E. Fortin, to be published.
[4] J. Berger, J. Castaing and M. Fisher, J. de Physique 40, 13 (1979).
[5] N. Caswell, J. S. Weiner and P. Y. Yu, Solid State Comm. 40, 843
(1981).
[6] E. Fortin, S. Fafard and A. Mysyrowicz, Phys. Rev. Lett. 70, 3951
(1993).
28
On the Bose-Einstein Condensation of
Excitons: Finite-lifetime Composite
Bosons
Sergei G. Tikhodeev
Institute of General Physics
Moscow 117333
Russia
Abstract
A brief review of experimental and theoretical work on Bose-Einstein con-
densation (BEC) in nonequilibrium systems is presented, with emphasis on
excitons in a semiconductor. Conditions for obtaining BEC in such sys-
tems are discussed, with special attention being paid to the effects of finite
lifetime, the role of spatial inhomogeneity and phonon-driven transport of
excitons.
1 Introduction
After many years of study, Bose-Einstein condensation (BEC) in a dilute
Bose gas has not yet been unambiguously established experimentally in
any real physical system. A systematic search has been undertaken in
atomic systems (atomic spin-polarized in magnetic field hydrogen [1]
and more recently in cesium [2]) and in excitonic systems in solids (see
e.g. [3, 4]).
Observation of BEC in an excitonic system is made easier [5, 6] because
of the following reasons: (i) small particle masses (of the order of the
free electron mass mo); (ii) the possibility of reaching sufficiently high gas
densities with an increase of excitation intensity; and (iii) the presence
of luminescence, making it possible to determine the particle distribution
function. However, excitons are finite-lifetime particles, and a number
of problems arise due to this fact. The goal of the present paper is
to analyze the effect of the finite lifetime on the possibility of BEC of
excitons.
In the case of incoherent excitation [7], a Bose-Einstein (non-
Maxwellian) distribution has been experimentally demonstrated in a sys-
524
On the BEC of Excitons: Finite-Lifetime Composite Bosons 525
n T
(cm-3) (K)
where
is the BEC temperature of an ideal Bose gas with density n, and nc(T) is
the critical density at temperature T. For comparison, the corresponding
numbers are shown for atomic hydrogen and excitonic systems. The
logarithm of 8 is the dimensionless distance from a point (n, T) on a
phase diagram to the BEC phase boundary given by (2); if S > 1, the
conditions of BEC are met. (The mysterious number 13 = 22 + 32 is the
sum of the squared numerator and denominator of the exponent 2/3 in
the BEC criterion (2) .) One can see from Table 1 that, up to now, the
search for BEC in excitonic systems has been in the lead, and, in the case
of paraexcitons in CU2O, the conditions for BEC in an ideal Bose gas
have been met [13].
526 S. G. Tikhodeev
2 Experiments under Stationary Excitation
Stationary excitation was used in earlier experiments in CU2O [9] and in
Ge [8]. The flow of thermalizing excitons should compensate the decay
of condensate particles. The latter introduces some deviations of BEC
conditions from the ideal Bose gas given by (2) [14, 15]. The physical
reason for these modifications of the phase diagram is that during its
lifetime, a newly generated exciton should have enough time to diminish
its energy via exciton-exciton and exciton-phonon collisions and to reach
the k = 0 state.
To illustrate the modifications of BEC conditions, a standard kinetic
equation for excitons interacting with a phonon thermostat at T — 0 was
solved in Refs. [14, 15]. In the statistical limit V —• 00, N/V —• n = const,
this equation takes the form
(3)
/
p<q
0<q<p
(4)
5 10 15 20 25 30 35
t(ns)
Fig. 1. Supersonic transport of orthoexcitons in CU2O at high power. The
distance A is measured from the surface to the innermost half maximum of
the spatial distribution, for the orthoexcitons as a function of time for several
laser-power densities. The dashed line has a slope given by the velocity of sound
S (from [11]).
125-
100 :
-15 5 15 25
t(10" 9 s)
Fig. 2. The time dependence of the FWHM of the carrier concentration for
different values of dimensionless excitation energy a = E/Es (Eq. (7)) in the case
of surface excitation (from [23]).
the FWHM of the exciton density on time and excitation intensity, which
agrees well with the experimental data. One can see in Fig. 2 that if
mhcoS2Ti
E>ES = (7)
References
[I] LF. Silvera and M. Reynolds, J. Low Temp. Phys. 87, 343 (1992).
[2] E. Tiesinga, A.J. Moerdijk, BJ. Verhaar and H.T.C. Stoof, Phys.
Rev. A 46, 1167(1992).
[3] E. Hanamura and H. Haug, Phys. Rep. 33C, 210 (1977).
[4] M. Ueta, H. Kanzaki, K. Kobayashi, Y. Toyozawa, and E. Hana-
mura, Excitonic Processes in Solids (Springer, Berlin, 1986).
[5] S.A.Moskalenko, Fiz. Tv. Tela 4, 276 (1962).
[6] J.M. Blatt, K.W. Boer, and W. Brandt, Phys. Rev. 126, 1691 (1962).
[7] We do not consider here the experiments with direct excitation of
coherent excitonic and biexcitonic states by an external coherent
source (EM wave) (see e.g. Refs. [3], [4] and M. Hasuo, N. Naga-
sawa, T. Itoh, and A. Mysyrowicz, Phys. Rev. Lett. 70, 1303 (1993)).
[8] I.V. Kukushkin, V.D. Kulakovskii, and V.B. Timofeev, JETP Lett.
34, 36 (1981).
[9] D. Hulin, A. Mysyrowicz, and G. Benoit a la Guillaume, Phys. Rev.
Lett. 45, 1970 (1980).
[10] D.W. Snoke, J.P. Wolfe, and A. Mysyrowicz, Phys. Rev. Lett. 59,
827 (1987).
[II] D.W. Snoke, J.R Wolfe, and A. Mysyrowicz, Phys. Rev. B 41, 11171
(1990).
[12] D.W. Snoke, Jia Ling Lin, and J.R Wolfe, Phys. Rev. 43, 1226
(1991).
[13] Further results for CU2O are reported in this volume. J.L. Lin
and J.R Wolfe obtain a Silvera-Reynolds number S = 2.8 (for
paraexcitons in stressed CU2O), and E. Fortin, E. Benson, and
A. Mysyrowicz report d as large as 5.
[14] S.G. Tikhodeev, Solid State Commun. 72, 1075 (1989).
[15] S.G. Tikhodeev, Sov. Phys. - JETP 70, 380 (1990).
[16] H. Frohlich, Int. J. Quantum Chem. 2, 641 (1968).
[17] N.G. Duffield, J. Phys. A: Math. Gen. 21, 625 (1988).
[18] D. Snoke and J.R Wolfe, Phys. Rev. B 39, 4030 (1989).
[19] Yu. Kagan, B.V. Svistunov, and G.V. Shlyapnikov, Zh. Eksp. Teor.
Fiz. 101, 528 (1992).
[20] H.T.C. Stoof, Phys. Rev. A 45, 8398 (1992).
[21] E. Levich and V. Yakhot, Phys. Rev. B 15, 243 (1977).
On the BEC of Excitons: Finite-Lifetime Composite Bosons 531
[22] L.V. Keldysh, in Problems of Theoretical Physics, A Memorial Volume
to Igor E. Tamm (Nauka, Moscow, 1972), p. 433.
[23] A.E. Bulatov and S.G. Tikhodeev, Phys. Rev. B 46, 15053 (1992).
[24] J.R Wolfe, J. Lumin. 30, 82 (1985).
[25] S.G. Tikhodeev, Sov. Phys. Usp. 28, 1 (1985).
[26] L.V. Keldysh and N.N. Sibeldin, in Nonequilibrium Phonons in Non-
metallic Crystals, Eisenmenger and Kaplyanskii, eds. (Elsevier, Am-
sterdam, 1986), p.455.
[27] N.N. Sibeldin, V.B. Stopachinskii, S.G. Tikhodeev, and V.A. Tsetkov,
JETP Lett. 38, 207 (1983).
[28] V.V. Konopatskii, G.A. Kopelevich, and S.G. Tikhodeev, Zh. Eksp.
Teor. Fiz., in press (1994).
[29] B. Link and G. Baym, Phys. Rev. Lett. 69, 2959 (1992).
29
Charged Bosons in Quantum
Heterostructures
L. D. Shvartsman and J. E. Golub
Racah Institute of Physics
The Hebrew University of Jerusalem
Jerusalem 91904
Israel
Abstract
532
Charged Bosons in Quantum Heterostructures 533
electrostatic interaction of two carriers when one of them is near the
bottom of one band and another one near the top of the next band. In
such a case, the pairs are excitonic charged complexes. The other variant
corresponds to the case of two carriers near the top of the band, so that
both of them have a negative mass. In both situations, it is necessary
to stress that we are considering an excited state of the system. As will
be clear in the first case, a boson may be formed only when the reduced
mass of the pair is negative and so its total mass is positive. For two
negative-mass carriers, both the total mass and the reduced mass of the
pair will be negative.
In order to make clear the physical nature of the pairing mechanism,
let us write down a regular hydrogen-like Schrodinger equation :
[-(ft2/2m*)V2 + e<j>(p)]ip(p) = Exp(p) (1)
It does not matter if it is a three- or a two-dimensional equation. In
the usual case, when m* > 0 is the reduced mass of the pair, and <t>(p)
is an attractive electrostatic potential, (1) describes bound states with a
discrete spectrum. In the three-dimensional situation the potential well,
of course, is supposed to be deep enough, and in the two-dimensional
situation, as is well known, at least one confined level always exists.
It is obvious that if we put in Eq. (1) a negative reduced mass but
a repulsive potential $(p), we still have the same equation. This would
describe a bound state of the complex not in the well but in the barrier.
The kinetic energy of the internal motion has a negative value in this
case.
The idea of a charged, exciton-like complex appeared as early as
1971 in connection with the observation of a reversed hydrogen-like
absorption spectrum in bulk crystals of BH3 [7, 8]. The reversal of the
usual hydrogenic spectrum for excitons was explained as a result of the
existence of a bound complex of two electrons of different bands, with
negative exciton reduced mass. A similar effect has been reported in bulk
ZnP2 [9]. The complex of two electrons was called a bielectron, and the
complex of two holes was called a bihole.
In the present paper, we show that quantum heterostructures of cubic
semiconductors may exhibit biholes, and that this characteristic is a
general one, possible in nearly all quantum wells of cubic semiconductors.
Specifically, we show that two holes may form a bound pair, and compute
the binding energy including the effects of the two-dimensional screening.
Being a quasistationary state, the pairs have a finite lifetime. We conclude
that experimentally observable binding energies are realizable, and we
534 L. D. Shvartsman and J. E. Golub
8 -5
Fig. 1. Valence subband structure for a GaAs quantum layer in the approximation
of an infinite potential barrier, K = kL/n is the dimensionless wavevector, where
k is the usual wavevector and L is the well width. The energy e is measured in
units of h2n2/2moL2.
argue that the pairs are long-lived objects. A detailed calculation of the
lifetime is postponed for a later publication.
The possibility of charged bosons in heterostructures follows from
the form of the valence subband structure of quantum wells of most
cubic semiconductors. A generic band structure is shown in Fig. 1.
A general feature of the dispersion is the presence of negative mass
states. These states exist in the first excited subband of two-dimensional
films of almost all cubic semiconductors [10-12]. They appear as a
sequence of multicomponent anisotropic structure of the Kohn-Luttinger
hamiltonian. For example [10-12], in deep potential wells of some
materials (e.g., GaAs, Ge, GaSb, InSb, InP, InAs), this subband is the
n = 2 heavy hole band [10]. In others (e.g., AlSb, ZnTe, GaP), it is the
n = 1 light hole band [10]. In shallow wells, the band ordering may be
reversed. However, in all these systems, the sign of the effective mass is
negative [13].
Consider the electrostatic interaction between two holes when one
belongs to the ground subband (mass m\ > 0), and the other to the
excited subband (mass mi < 0). If the reduced mass (m*)"1 = mj"1 + m^1
is negative, the electrostatic repulsion, as we saw above, is converted to
a net attraction. It leads to the formation of a bound pair, that is, to
the formation of a bihole. A necessary condition for bihole formation
Charged Bosons in Quantum Heterostructures 535
1 2 3 4
x (percent)
is m* < 0 or
m2 (2)
Eq. (2) is satisfied in infinite quantum wells of all the materials mentioned
above except InSb and InAs. Eq. (2) may also be satisfied in finite wells
of properly chosen width and material composition. So, the possibility
of biholes is a general feature of quantum wells.
Bihole formation in quantum wells was first considered in 1983 [14].
We discuss below several important differences between experiments in
bulk crystals and the present work.
We next calculate the bihole binding energy using a quantum well
of AlGaAs/ InxGai_xAs/AlGaAs as a model system starting from a
Kohn-Luttinger multicomponent formulation. In any heterostructure,
changes in the doping x affect both the well-depth and the uniaxial
strain. The present system has the important calculational advantage
that the uniaxial strain dominates [15]. The calculation thus pro-
ceeds from a subband structure obtained assuming an infinite well, but
takes into account the x-dependent uniaxial strain in the well layer.
In particular, we assume the dependence of the effective masses upon
the indium mole fraction x shown in Fig. 2 [15, 16]. This assump-
tion permits us to calculate an x-dependent binding energy and then
adjust x to maximize the effect within the applicability range of the
model.
536 L. D. Shvartsman and J. E. Golub
Now let us consider (1) in the two-dimensional case. The electrostatic
potential c/>(p) in the presence of screening by the positive mass holes of
&keiKP(t>(K) (3)
the ground subband is given by
where [17-19]
c/>(K) = 2ne/e1(K + 2/a). (4)
Here, e = €2/e\; and ei(e2) is the dielectric constant of the well (barrier)
layers. In the numerical solutions below, we assume e = 1. Other
cases, for example, that of a free standing film, will be treated in a later
publication, a = eih2/m\e2 is a Bohr radius defined for holes in the first
subband. For p > L, we find for <f)(p) the asymptotic behavior
e P
/ '
Combining (7) and (8), we find for the x-dependent bihole binding energy
Charged Bosons in Quantum Heterostructures 537
References
[I] P.L. Gourley and J.P. Wolfe, Phys. Rev. Lett. 40, 526 (1978);
J. P. Wolfe and C. D. Jeffries, in Electron-Hole Droplets in Semi-
conductors, CD. Jeffries and L.V. Keldysh, eds. (North Holland,
Amsterdam, 1983) p. 431.
[2] D. Snoke, J.L. Lin, and J.P. Wolfe, Phys. Rev. B 43, 1226 (1991).
[3] D. Snoke and J.P. Wolfe, Phys. Rev. B 42, 7876 (1990).
[4] D. Snoke, J.P. Wolfe, and A. Mysyrowicz, Phys. Rev. Lett. 64, 2563
(1990).
[5] I.F. Silvera, Physica 109+110B, 1499 (1982); T. J. Greytak and D.
Kleppner, in New Trends in Atomic Physics, G. Grynberg and R.
Stora, eds. (North Holland, Amsterdam, 1984), Vol. II.
[6] E.L. Raab, M. Prentiss, Alex Cable, Steven Chu, and D. E. Pritchard,
Phys. Rev. Lett. 59, 2631 (1987), and references therein.
[7] E.F. Gross, V.I. Perel, and R.I. Shechmamet'ev, JETP Lett. 13, 229
(1971).
[8] E.F. Gross, N.V. Starostin, M.P. Shepilov and R. I. Shechmamet'ev,
Proceedings of the USSR Academy of Sciences 37, 885 (1973).
[9] A.V. Selkin, I.G. Stamov, N.U. Syrbu, and A.G. Umenets, JETP
Lett. 35, 57 (1982).
[10] L.D. Shvartsman, Ph.D. thesis, USSR Academy of Science Siberian
Branch, Institute of Semiconductors, Novosibirsk (1984).
[II] A.V. Chaplik and L.D. Shvartsman, Poverkhnost (Surface), No. 2,
73 (1982).
[12] L.D. Shvartsman, Sol. State Commun. 46, 787 (1983).
[13] Silicon is the only exception to this rule. See Refs. [10] and [12].
540 L. D. Shvartsman and J. E. Golub
[14] A.V. Chaplik and L.D. Shvartsman, in All Union School on Surface
Physics, (USSR Academy of Sciences, Chernogolovka, 1983), pro-
ceedings of the Third All Union School on Surface Physics held in
Tashkent, October, 1983, p. 123.
[15] B. Laikhtman, R.A. Kiehl and DJ. Frank, J. Appl. Phys. 70, 1531
(1991).
[16] O.V. Kibis and L.D. Shvartsman, Poverkhnost (Surface), No. 7, 119
(1985), in Russian.
[17] N.S. Rytova, Vestnik Moskovskogo Universiteta (Moscow University
Proceedings), No. 3, 30 (1967).
[18] L.V. Keldysh, JETP Lett. 29, 658 (1979).
[19] A.V. Chaplik and M.V. Entin, JETP 34, 1335 (1972).
[20] G. Bastard, in Physics and Applications of Quantum Wells and Su-
perlattices, E.E. Mendez and K. von Klitzing, eds. (Plenum, New
York, 1987).
[21] Direct decay can always be made forbidden by slight changes in
material composition, or other quantum-well parameters.
30
Evidence for Bipolaronic Bose-liquid and
BEC in High-Tc Oxides
A. S. Alexandrov
Interdisciplinary Research Centre in Superconductivity
University of Cambridge
Madingley Road
Cambridge, CBS OHE
UK
Abstract
Recent experiments on the near-infrared absorption, thermal conductivity
and the critical field HC2 in several high-Tc oxides are interpreted as a
manifestation of the Bose-Einstein condensation of small bipolarons.
541
542 A. S. Alexandrov
conducting transition on the near-infrared absorption [5], the thermal
conductivity enhancement below Tc [6] and the upper critical field, di-
vergent at low temperature [7]. These observations can be interpreted
as a manifestation of the Bose-Einstein condensation of bipolarons in
copper-based oxides.
and fi = 0 if T < Tc. Here n is the total number of pairs per cell and
the density of states is assumed to be energy independent within bands.
Evidence for Bipolaronic Bose-liquid and BEC in High-Tc 543
1.2 I I I I
1.0 —
0.8 -
nfi I I I
T/Tc
r x dx
* <*
Jo
where K^o is temperature independent, t = T/Tc is the reduced temper-
ature and rj is proportional to the charged impurity concentration.
With Eq. (5) one can explain the experimentally observed enhancement
of the thermal conductivity below Tc, shown in Fig. 2. It should be
mentioned that the commonly accepted explanation of this enhancement
due to the lattice contribution to the heat transport has now been
rejected [11].
For carriers with a double elementary charge, the Lorentz number
should be at least four times smaller than for degenerate electrons. This
fact enables one to understand [2] the near equality of the in-plane
thermal conductivity above 100K in the insulating and 90K crystals of
YBa 2 Cu 3 0 7 -a [12].
30
r\ =0.01
a/ '••
1.6 - 2^ 25
E 7/'b\'-. pre - an
/ \^ • post - an
\ ^2°
\«\\
/ 0.02
2 1.2 -
\ •AN
o
1 / \ ^nr.-.r.:-.-.:.: ~
I / 0.05
\ ^ \ / ,40
, \ ;
"I 0.8 80 120 160
T(K)
N\
0.4 - ^ ^
where 2e is the charge of a boson; A(r), l/imp(r) and ft are the vector,
random, and chemical potentials, respectively.
The definition of H* in Eq. (6) is identical to that of the upper
critical field H& of BCS superconductors of the second kind. Therefore
H* determines the upper critical field of bipolaronic or any "bosonic"
superconductor.
The first non-trivial delocalized solution of Eq. (6) appears at \i = Ec.
546 A. S. Alexandrov
Thus the critical curve H*(T) is determined from the conservation of the
number of particles n under the condition that the chemical potential
coincides with the mobility edge Ec (for more details see Ref. [7]):
/c=0
WlW
*-7. <io»
jVL(6) = ^ e x p ( ^ — ^ ) , (11)
1.0
CD
0.8
• \ \ \
Kv
\ \ \
\ n L/n = o
0.6
a4
V \
.1 0.4 . \ \08
I6 0.2
/—^x.
1.1
v° \
\.
0.0
0.0 0.2 0.4 0.6 0.8 1.0
Temperature T / T co
5 Conclusion
Earlier evidence for small (bi)polarons in copper-based oxides comes
from the photoinduced infrared absorption, measured by Heeger and by
Taliany [18, 19], from the observation by Sugai [20] of both infrared and
Raman-active vibration modes, and from XAFS results on the radial
distribution function of apex oxygen ions [21, 22].
The experimental observations discussed above, and many other fea-
tures of high-Tc copper oxides [3, 4], lead us to the conclusion that a
charged Bose-liquid of small intersite bipolarons is a simple but far-
reaching model of high- Tc superconductors.
Abstract
We study the dynamic structure function S(k,co) of Bose liquids in the
asymptotic limit k,co —• oo at constant y = j(co — k2/2m), using the or-
thogonal correlated basis of Feynman phonon states.
1 Introduction
In the last few years there has been a growing interest in the possibilities
for experimental determination of single-particle momentum distributions
in non-relativistic many-body systems by means of inelastic neutron
scattering at large momentum transfers. This interest was generated by
the advent of pulsed-neutron sources that made possible measurements
with substantially larger momentum transfers then before. For example,
in the case of liquid 4 He a momentum transfer of 20-30 A" 1 was
achieved [1], which is much larger than the rms momentum of atoms in
the liquid of ~ 1.6 A" 1 . If it is assumed that at such large momentum
transfers, the potentials between the atoms in the liquid are negligible
compared to the large kinetic energy of the struck atom, the deep inelastic
response is completely determined by the initial single-particle momentum
distribution. This assumption is known as the impulse approximation
(IA) [2, 3]. At large momentum transfers fe, the dynamic structure
function (DSF) 5(k,co) exhibits the phenomenon of y-scaling, i.e. the
combination J(y) = ^S(k,co), called the scaling function, depends solely
upon the scaling variable [2] y = f(co — ^ ) and not separately upon k
and co. The scaling function in the IA is given by:
i r00
dqq
JiA(y) = ncS(y) + ^ 5 " I
550
The Dynamic Structure Function of Bose Liquids 551
where p is the density of the system, n(q) is the single-particle momentum
distribution and nc the corresponding condensate fraction. However, the
^-scaling property of the deep inelastic response is more general than the
IA in the sense that it may hold even when the IA fails.
If the IA is valid, the momentum distribution can be extracted from
the deep inelastic response in a model-independent way. This gives rise
to the exciting possibility that the momentum distribution is a directly
observable quantity. If this were true, not only would we be able to check
the accuracy of our calculations and determine interparticle potentials
more reliably, but also we could test fundamental physical ideas such as
Bose condensation!
Unfortunately, in the most interesting cases, such as liquid 4He, there
is a strong repulsive short-ranged component in the interparticle po-
tential which is not negligible compared to the recoil kinetic energy
even for the large momentum transfers achieved in recent experiments.
Thus the IA must be corrected to account for interaction of the re-
coiling particle with the surrounding medium. These corrections are
called final state interaction (FSI) corrections. Clearly, it would be
desirable to develop an analytic, first-principles theory of the dynamic
structure function at large momentum transfers, in order to understand
the FSI.
There were many attempts that start from the IA but take into account
the interactions in the initial state of the target, and improve upon it
by incorporating the effects of FSI [4]. This procedure becomes more
and more difficult as k decreases and the FSI become more important.
In contrast the response at small k and co is easily treated by using
Feynman phonon states [5]. In particular the orthogonal correlated basis
(OCB) formalism [6] based on Feynman's ideas is quite successful [7]
in explaining the observed S(k,co) at k < 2 A'1. The small and large k
methods have different starting points; the momentum distribution n(q)
is the main input for the IA, while the OCB formalism uses the static
structure function S(q) as the main input.
Although in principle the OCB formalism can be used to study S(k,co)
at all k and co, in practice the calculations [7] become technically com-
plicated as k increases. Nevertheless we have shown [8] that, by using
field-theoretical techniques in OCB perturbation theory, it is possible
to calculate the S(k,co) in the scaling limit, and thus extend this low
k method to the k —> oo limit. In this paper, we present results of
this approach as applied to the case of a Bose liquid at zero tem-
perature.
552 A. Belie
2 OCB Calculation of S(k,co) in the Scaling Limit
The DSF S(k,co) can be written as
1
S(k,co) = —ImD(k, co), (2)
n
where
k
N
n=0
where
The Dynamic Structure Function of Bose Liquids 553
and | 0) is the ground state of the system. FP states do contain the
appropriate short-range correlations, and are reasonably close to the
exact eigenstates of the Hamiltonian. Thus we expect Hr to be small
and series (3) to be rapidly convergent. However, they are not mutually
orthogonal, which is the general property of correlated basis (CB) states.
Such states can be used in the perturbation expansions, but require a
special formalism that takes into the account their non-orthogonality, and
it is not clear whether the non-orthogonality effects that are introduced
when the expansion is truncated are negligible or not. In order to use
normal perturbation theory, we have to orthogonalize the CB states.
This can be achieved using the Lowdin transformation [10], but resulting
OCB states have higher energies than CB states. Perturbative corrections
move the energy down again, and both the increase in energy due to
the Lowdin transform and the decrease due to perturbative corrections
are larger than the net displacement [6]. In view of these difficulties, the
OCB perturbation theory (OCBPT) is not the preferred tool for studies
of quantum liquids. Recently, however, a new orthogonalization scheme,
free of the problems mentioned above, was proposed [11]. It is expected
that the convergence of OCBPT does not depend upon the specific nature
of the bare interaction, i.e. if it has a hard-core or not. For these reasons,
we use it to study the scaling properties of the response at large k and co.
When the orthogonalized FP in (6) are used as intermediate states in
(3), it can be shown [8] that the diagonal matrix elements HQJJ are of
order O((fe/m)2), while the off-diagonal matrix elements H[j are of order
0(k/m\ from which the scaling property
-1 >, (9)
The function /(/) depends on the static structure function S(q) only, and
is given by:
(11)
_n=0
where the coefficients ocn = 1,-1/2,3/8,... are those that appear in the
expansion of (1 + *)~ 1/2 , and the functions /„(/) are defined recursively
as:
(12)
r d3h |
), n>\. (13)
Eqs. (9)—(13) represent the main results of this work.
A „ r , r / 1 sin(s/)
r° ,,sin(s -i ,
A = AnpC I ds exp —?- / dl—V"
me
B = Cy. (16)
The scaling function in IA (Eq. (1)) exhibits the same small y behavior
with the coefficients
f00 me
CIA = nc, AIA= / dqqn(q), BIA = limqn(q) = nc— 9 (17)
2
Jo «^°
and it follows that the ratio B/C in our theory is identical to that in IA:
£ =^. (18)
C CIA
Next, we want to discuss the origin of the S -function peak at y = 0. In
general, the presence of the ^-function in DSF means that probe couples
to a long-lived state with a given energy and momentum. The inverse
lifetime of this excitation is expected to be proportional to k on the basis
of the semi-classical value T" 1 ~ pva, where v = k/m is the velocity
and a the average two-body scattering cross section. However, at y = 0
this argument fails, as can be seen in the second-order contribution
to (3). In that order, an off-shell FP with momentum k and energy
co = k2/2m + ky/m can decay into two FP of momenta k — 1 and 1, thus
acquiring a lifetime
o< 0.3 -
Fig. 1. The scaling functions of liquid 4He at equilibrium density and zero
temperature. Solid line: J(y) of Eq. (9) obtained using S(q) of Ref. 13; dotted
line: JiA{y) generated from n(q) of Ref. 14. The ^-function peaks at y = 0, with
the respective strengths of 0.15 and 0.092, are not shown.
are satisfied by J(y) in (9). For the y2 (kinetic energy) sum-rule, we obtain
<} /»OO 1 /*OO
MD = (22)
and that the scaling function J(y) of (9)-(10) reduces to (1), as expected.
When (9)—(13) are applied to the case of liquid 4 He, using the mea-
sured [13] S(q), the value C = 0.15 and the J(y) shown in Fig. 1 are
obtained. For comparison, JiAiy) generated from the variational n(q)
(Ref. [14]) which has the condensate fraction nc = 0.092, is also shown.
Finally, using the value Ek = 14.8K from Ref. [14] and S(q) from
Ref. [13], we find that
3 r
* dll2f(l) = 0.91. (23)
As was argued above, the closeness of this value to 1 suggests that the
The Dynamic Structure Function of Bose Liquids 557
neglected terms in expansion (3) are much smaller than those taken into
account.
References
[I] RE. Sokol, in Momentum Distributions, R.N. Silver and RE. Sokol,
eds. (Plenum, New York, 1989).
[2] G.B. West, Phys. Rep. 18C, 263 (1975).
[3] RC. Hohenberg and P.M. Platzman, Phys. Rev. 152, 198 (1966).
[4] R.N. Silver, in Momentum Distributions, R.N. Silver and RE. Sokol,
eds. (Plenum, New York, 1989); A.S. Rinat, Phys. Rev. B 40, 6625
(1989).
[5] R.R Feynman, Phys. Rev. 94, 262 (1954).
[6] E. Feenberg, Theory of Quantum Liquids (Academic Press, New
York, 1969).
[7] E. Manousakis and V.R. Pandharipande, Phys. Rev. B 33,150 (1986).
[8] A. Belie and V.R. Pandharipande, Phys. Rev. B 45, 839 (1992).
[9] H.A. Gersch and L.J. Rodriguez, Phys. Rev. A 8, 905 (1973).
[10] P.O. Lowdin, J. Chem. Phys. 18, 365 (1950).
[II] S. Fantoni and V.R. Pandharipande, Phys. Rev. C 37, 1697 (1988).
[12] A. Belie and V.R. Pandharipande, Phys. Rev. B 39, 2696 (1989).
[13] H.N. Robkoff and R.B. Hallock, Phys. Rev. B 25, 1572 (1982).
[14] E. Manousakis, V.R. Pandharipande, and Q.N. Usmani, Phys. Rev.
B 31, 7022 (1985); ibid. B 43, 13587 (1991).
32
Possibilities for BEC of Positronium
P. M. Platzman and A. P. Mills, Jr
AT&T Bell Laboratories
600 Mountain Avenue
Murray Hill, New Jersey 07974
USA
Abstract
We review the proposal for Bose-Einstein condensation of positronium
atoms. All of the ingredients necessary to achieve BEC of Ps atoms are
currently available.
1 Introduction
In this volume several authors have discussed and described a variety
of weakly interacting systems which might display BEC. In this short
contribution we suggest that a dense gas of positronium (Ps) atoms in
vacuum is a rather ideal but somewhat more exotic system that might
be a very good candidate for observing a weakly interacting BEC.
Recent investigations of the interactions of positrons (e+) and (Ps)
with solids have led to extraordinary improvements in the kinds of low
energy experiments we can do with the positron [1, 2]. All of the
ingredients necessary to achieve BEC of Ps atoms are currently available.
We envision a scenario where roughly N = 105 Ps atoms are trapped in a
volume V ~ 10~13 cm3 and allowed to cool through the Bose transition
temperature of 20 — 30K in a time of the order of nanoseconds. In the
following we discuss the relevant interactions and describe how Ps BEC
can be achieved.
558
Possibilities for BEC of Positronium 559
is half that of H. The ground state of Ps is a spin singlet separated from
an excited triplet by an energy AEST — 10K. Ps annihilates itself, i.e., it
becomes high energy 7-rays [3],
The annihilation characteristics of the Ps atom are dependent on which
of the two ground states it is in. The ground state singlet (s) is short
lived with a lifetime TS = 1.25 x 10~10 s. For a Ps atom at rest, the decay
occurs with the emission of two 0.5 MeV y-rays which come out precisely
in opposite directions due to momentum conservation. If the Ps atom
is moving with momentum /?, the y -rays come out with a small angle
0 ^ p/mc relative to each other. The nearby triplet state (t) is prohibited
by selection rules from decay into the two y -ray channel. Instead it decays
by three y -rays with a spread of energies and a much longer lifetime of
Tt = 1.42 x 10~7 s. A magnetic field mixes the triplet with singlet, thus
rapidly quenching the triplet.
Like hydrogen, the PS2 molecule exists in an overall singlet state [4]. It
has a binding energy EM = 0.4 eV. Two low energy Ps atoms scatter from
one another with a cross section o = Ana1 determined by how close the
bound state is to the continuum [5]. Specifically, as = (2mEM)~{^2 = 3 A,
and GS = 10~14 cm2. On the other hand the non-singlet channel has
a scattering length more like a Bohr radius, i.e., at = 1 A and ot =
10~15 cm2. Of course, the experiment we are considering here would,
for the first time, be sensitive to such cross sections and a measurement
of them should be possible.
Vacuum Solid
Polarized
e+(=10keV) Plasmon
•HOeV)
Diffusion
Thermal e+
Fig. 1. Schematic of the interaction of a 10 keV beam with a typical solid such
as Si.
4 Dense Positronium
The "initial" condition for the Ps condensation experiment is sketched
in Fig. 2. A bunched 1 nanosecond brightness-enhanced microbeam
(diameter ljum) consisting of N > 106e+ at an energy of 5 keV is incident
at time t = 0 (1 nanosecond uncertainty) on the surface of a Si target
which has, at the point of entry, a small cavity etched into it. We
choose Si rather than a metal for several reasons, to be discussed shortly.
The geometry we envision will typically be a cylindrical cavity of 1 fim
diameter with a height of 1000 A. These dimensions are rather easily
achieved with conventional lithography techniques. The e+9 as discussed,
stop in the Si at a depth of about 1000 A. About one-quarter of the
initial e+, because of diffusion and the negative work function for Ps, will
be reemitted as electron volt Ps into the cavity. The singlet Ps rapidly
decays leaving us with a hot gas of bosons (triplet Ps atoms), at a density
n = N/V ^ 1018 cm"3.
In Ref. [6] we consider in some detail the time evolution of this
confined hot gas of Ps atoms due to collisions with the wall and with
other Ps atoms in the cavity. We conclude that such a gas would indeed
cool to a Bose condensed state by phonon emission at the walls in a
time short compared to a triplet lifetime. However, the serious difficulty
which must be avoided is quenching of the triplet state by collisions of
Ps atoms with the same wall and with each other.
Scattering from the walls occurs at a very rapid rate, with Tw — 1012
1
s" for Ps atoms with a energy of 0.1 eV. If the solid wall of the cavity
is metallic, i.e., there are unpaired spins, then exchange of the e~ in
the triplet Ps with e~ in the metal occurs with high probability and the
triplet state is quenched very rapidly. If the wall is insulating, i.e., the
electron spins are paired, then exchange scattering is often energetically
562 P. M. Platzmanand A. P. Mills, Jr
1 st remoderator
10 6 e+
1cm
5nsec
1mm
30jim|[
forbidden and the only allowed process is a weak relativistic effect, which
is negligible.
The scattering of triplet Ps atoms with one another is the mechanism
by which the system equilibrates internally. This type of scattering can
also lead to exchange scattering which will annihilate the gas before it
reaches BEC. In the s wave scattering approximation the bulk triplet-
triplet scattering rate characterized by an s-wave scattering length of
lA is roughly Tt = 1010 s"1. It is rapid enough to equilibrate the gas as
it cools.
In Ref. [6] we have shown that collisions between triplets with their
spin pointing in different directions will surely lead to conversion of
triplet to singlet at roughly the same rate. However, it is possible to
prevent this type of annihilation by using a fully polarized e+ beam.
Fully polarized e+ beams are commonly available since all beta-decay
sources generate e+s by a parity non-conserving process which leads to
e+ polarized with their spin in the direction of their momentum. Slowing
down inside a low Z solid such as Si leaves the initially fully polarized
e+s about 50% percent polarized (see Berko in Ref. [2]).
For such a polarized beam of e+ incident on an unpolarized e~ target,
it is easy to show, based on rather general arguments [6], that triplet-
Possibilities for BEC of Positronium 563
Counts'.
Bose
condensate
Maxwellian
Ps
e
Fig. 3. Sketch of a hypothetical y-ray angular correlation spectrum for BEC Ps.
The double humped structure is characteristic of the Bose condensed fraction.
References
[1] A.P. Mills, Jr, in Positron Solid State Physics, edited by W. Brandt
and A. Dupasquier (North Holland, Amsterdam, 1983) p. 421; P.J.
Schultz and K.G. Lynn, Rev. Mod. Phys. 60, 701 (1988).
[2] Positron Studies of Solids Surfaces and Atoms, edited by A.P. Mills,
Jr, W.S. Crane and K.F. Canter (World Scientific, Singapore, 1985).
[3] H. Bethe and E. Salpeter, Quantum Mechanics of One and Two Elec-
tron Atoms (Academic, New York, 1957).
[4] E.A. Hylleraas and A. Ore, Phys. Rev. 71, 493 (1947); Y. K. Ho,
Phys. Rev. A 33, 3584 (1986). For a review see M.A. Abdel-Raouf,
Fortschr. Phys. 36, 521 (1988).
[5] T.Y. Wu and T. Ohmura, Quantum Theory of Scattering (Prentice Hall,
New Jersey, 1962).
[6] P.M. Platzman and A.P. Mills, Jr, Phys. Rev. B 49, 454 (1994).
33
Bose-Einstein Condensation and Spin
Waves
R. Friedberg and T. D. Lee
Department of Physics
Columbia University
New York, NY 10027
USA
H. C. Ren
The Rockefeller University
New York, NY 10021
USA
The Quantum Lattice Model [1] is a natural basis for computer simu-
lations of Bose liquids as well as for possibly realistic models of mobile
bosonic excitations in a solid. It depicts particles obeying Bose statistics
occupying sites of a lattice and possessing kinetic energy by virtue of a
hopping term —2 ^2,{b\bj + b^bt) in the Hamiltonian, where ^ counts
each nearest neighbor pair once; but it is understood that no more than
one particle may occupy the same site.
It is of interest to add an attractive two-body potential at nearest
neighbor separation, so that the Hamiltonian may be written
bt + 2 ]T(bJ - b])(bt - bj) - a ^ b]b)btbj. (1)
565
566 R. Friedberg, T. D. Lee and H. C. Ren
can be calculated in powers of F, which itself remains finite as G —• +00.
Particularly important is the long-wavelength scattering amplitude
where go is the solution at the origin of the Poisson equation with unit
source. The continuum analog of |yo is 4na/m, where a,m are the
diameter and mass of the spheres.
If a 7^ 0 the same treatment can be applied with
1 4
" a
go ~ ^ (4)
If we define
so that
[ci,c}]=a*diJ = (l-2c)ci)5ij, (7)
cjc). (8)
A). (10)
The parameter 9 is chosen so that Hjjeis leads to a zero value of < erf > .
Then, by keeping only those terms of Hfjeis quadratic in Cj and cj,
they obtained an approximate spectrum of excitations in the presence
of the condensate. This spectrum is phonon-like at low momentum;
the phonons are the Goldstone bosons associated with the breaking of
azimuthal symmetry by setting 9 ^= 0.
We have made (2) the starting point of a systematic expansion by using
the connection between (2) and (4) a second time in reverse; that is, we
replace i/# ds by a new bosonic Hamiltonian H obtained by replacing
c\ by a new boson operator b\ and adding a term G ^ . 6fft,-&,-&,• where
G -> +00 [2]. Because of the 0-rotation, < ft, >= 0; 9, not < ft >, is
the order parameter for Bose condensation. The new Hamiltonian H
contains terms in btbj and ftjft; so that the ground state has nonzero
values of < ft, ft; > and < ftjft; >, etc. These and other quantities
(speed of sound, spectrum, ground state energy) can be calculated from
diagrams as a power series in A. For furthur details, see Ref. [2].
References
[1] T. Matsubara and H. Matsuda, Prog. Theor. Phys. 16, 569 (1956).
[2] R. Friedberg, T.D. Lee and H.C. Ren, Ann. Phys. (N. Y.) 228, 52
(1993).
[3] T.D. Lee and C.N. Yang, Phys. Rev. 105, 1119 (1957); 113, 1165
(1959).
[4] F. Dyson, Phys. Rev. 102, 1217, 1230 (1956).
34
Universal Behaviour within the
Nozieres-Schmitt-Rink Theory
F. Pistolesi and G. C. Strinati
Scuola Normale Superiore
1-56126 Pisa
Italy
Abstract
We show that the natural variable to follow the crossover from Cooper-pair-
based superconductivity to Bose-Einstein condensation within the model of
Nozieres and Schmitt-Rink is the product kp£>, where kp is the Fermi wave
vector and £, is the coherence length for two-electron correlation. In terms
of this product, the results of the model do not depend on the detailed
form of the (separable) pairing potential, and the crossover turns out to be
restricted to the universal region n~l< kp£, < 2n. Experimental estimates
indicate that kpt, « 10 (> 2%) for high-Tc superconductors.
569
570 F. Pistolesi and G. C. Strinati
H =
k,<7
a
/2,Ta -k+q/2,i a -k' + q/2,l^ +q /2,T
^,k'4+ q /2,T
k,k',q
where ako. is the destruction operator for fermions with wave vector k
and spin <r, ek is a single-particle dispersion relation, and Vk# is an
"effective" fermionic attraction. In its simplest form, ek = k2/2m* — fi,
where m* is an effective particle mass and \i is the chemical potential.
The variational calculation with the ground state BCS wave function
pairing potential:
(6)
2 = /<frg(r)r2 = EklVkffkl2
Jdrg(r) £ k l<Pk| 2 '
1.0
o
A
0.5
-1.0
10' 10°
Fig. 1. Chemical potential \i versus kF^ (at zero temperature). For the normal-
ization of pL and the meaning of the different curves see the text. The two limiting
curves corresponding to the values 10~5 and 104 of n/kl are marked with arrows.
The two curves shown for the reported extreme values of n/k$ m a v ^>e
regarded as limiting curves for all practical purposes. The remarkable
feature of Fig. 1 is that, when expressed in terms of kf£, the behavior
of the chemical potential is universal (i.e., independent of the parameter
fe0 of the interaction potential), but possibly for an "intermediate" range
rc"1- kF£ i 2n [7]. This universal behavior of \i versus kF£ suggests that
kF£ is the appropriate variable to follow the evolution from BCS to BEC.
Note also from Fig. 1 that \i is pinned to (about) the normal-state
value eF when kpt; > 2n, and that \i drops rather abruptly from eF at
kr€ ^ 2TT. Fig. 1 thus shows that, when the coherence length £ equals
the Fermi wavelength X? = 2n/kp, the system becomes unstable against
bosonization and the Fermi surface disappears. We expect that the
instability of the Cooper-pair-based superconductivity when kF£ ~ In
should persist beyond the limits of validity of the procedure we have
adopted to establish it. The stability criterion kF£ > In could then
be regarded as the analog of the Ioffe-Regel criterion for transport in
disordered systems.
One of us (F.R) gratefully acknowledges partial research support from
Europa Metalli-LMI S.p.A.
Universal Behaviour within the Nozieres-Schmitt-Rink Theory 573
References
[1] P. Nozieres and S. Schmitt-Rink, J. Low. Temp. Phys. 59, 195 (1985).
[2] A.J. Leggett, in Modern Trends in the Theory of Condensed Matter,
A. Pekalski and J. Przstawa, eds. (Springer-Verlag, Berlin, 1980), p.
13.
[3] S. Schmitt-Rink, C M . Varma, and A.E. Ruckenstein, Phys. Rev.
Lett. 63, 445 (1989); R. Friedberg and T.D. Lee, Phys. Rev. B 40,
6745 (1989); M. Randeria, J. Duan, and L. Shieh, Phys. Rev. Lett.
62, 981 (1989) and Phys. Rev. B 41, 327 (1990); L. Belkhir and M.
Randeria, Phys. Rev. B 45, 5087 (1992); M. Randeria, N. Trivedi,
A. Moreo and R.T. Scalettar, Phys. Rev. Lett. 69, 2001 (1992); R.
Cote and A. Griffin, Phys. Rev. B 48, 10404 (1993).
[4] Cf., e.g., J.R. Schrieffer, Theory of Superconductivity (Benjamin, New
York, 1964), Chapter 2.
[5] We have eliminated the Hartree-Fock-like terms by setting Fk,k = 0
for the diagonal components. This choice will not invalidate our
results, since these terms turn out to be irrelevant in the parameter
region of interest.
[6] NSR (Ref. [1]) state that r 0 ~ fcj^1 for G -> oo. Since bosonization
can be achieved only when r$ <C rT1, NSR are able to follow the
evolution from BCS to BEC as a function of G only in the "dilute
limit" n/feg <C 1 for the reduced (three-dimensional) density, that is,
for given density n only when fco > kp. This limitation prevented
NSR from connecting the two limits (BCS and BEC) irrespective of
fco. Our conclusion that ro ~ fcj^G"1/2, on the other hand, enables
us to satisfy the bosonization condition (W//CQ)/G 3 / 2 <C 1 even in the
"dense limit" n/k$ > 1, provided G is large enough.
[7] We have verified that the universal behavior of Fig. 1 for kf£ > 2n
and kft; < TT"1 is independent of the choice of the single-particle
dispersion relation 6k and of the shape of the interaction potential
Wk (barring pathological cases).
35
Bound States and Superfluidity in
Strongly Coupled Fermion Systems
G. Ropke
Fachbereich Physik
Universitat Rostock
Universitdtsplatz 3
D 18051 Rostock
Germany
Abstract
574
Bound States and Superfluidity in Fermion Systems 575
between pairs [9] (see also Ref. [10] for pairing in two dimensions). A
finite temperature Green function approach describing both two-particle
(Brueckner ladders) and pairing correlations has been considered in
Ref. [11], but a self-consistent solution of the gap equation has not been
obtained so far (see Zimmermann [3]).
Clearly one would like an improvement of the Hartree-Fock-Gorkov
mean-field description. As is well known, the Gorkov equation corre-
sponds to a Hartree-Fock approximation in the normal state. A thermo-
dynamic Green's function approach has been developed for the normal
state which is able to describe the formation of bound states and their
dissolution at higher densities (Mott effect, see Ref. [9]). We will give an
approach which includes both correlations and bound state formation in
the superfluid state. In this way, the equation of state is obtained for the
normal state and the superfluid state in a consistent way. Furthermore,
a unified description of BEC and BCS can be given, with a generalized
Gorkov equation.
The fermion system is described by the Hamiltonian
a
q-pi =
where
21"/
with
p
l
The anomalous (or off-diagonal) mean values {bibi)1 vanish.
Expressing H = H — J2i ^ i a ^ i i n the ^-representation, we have
><l+P2\>(l+ Pi t p ^ i
. (8)
Similar expressions are obtained for other spin combinations. In terms
of these new operators, in place of the original distribution function
given in (2)-(4), we can use the mean values < b\b\ >l, < b^b^b^br >l
to describe the system. We will not derive the kinetic equations for
the corresponding quantities N(l,t), 5(12,0, C(12,l/2/,t) (see Ref. [12]),
but will only discuss these expressions for thermal equilibrium. A sys-
tematic perturbative treatment can be given of the statistical operator
Bound States and Superfluidity in Fermion Systems 577
Z o * exp[—PH] with the help of the usual Matsubara Green function
technique. As pointed out above, we assume that only single-particle
and two-particle states are of relevance. In the cluster-Hartree-Fock
approximation (i.e., interactions involving the single-particle and two-
particle states in the medium are taken in the Born approximation), we
obtain the following diagrammatic representation for the single-particle
and two-particle Green functions (compare Ref. [14] for the normal state
case):
IiL
ex
ex
(9)
with
at
= lTn eJJ[
—00
s + i t n t / h
^ % " ^ " (13)
l'2', t)
VI!
7(13, 1'3') follows from 7(13,1'3'), with k T<-> -k |, p |<-> -/? |, and
E/€
two-particle energies lie well above the energy of the Bose condensate at
Econd(2q) = 2JU. Furthermore, the two-particle binding energy (difference
between the continuum edge of scattering states and the bound state
energy) depends on the total momentum and decreases with increasing
density. The Mott effect arises when the bound state energies merge with
the continuum of scattering states. This illustrates that our approach
to strongly coupled superfluidity corresponds, in the normal phase, not
only to a Hartree-Fock approximation, but also to a self-consistent
description of two-particle correlations [9], In particular, our present
approach includes the formation of bound states in a correlated medium
and their disappearance (or unbinding) at high densities. The Pauli
blocking is responsible not only for the dissolution of the bound states,
but it also arises as the essential mechanism to establish the superfluid
condensate in the strong-coupling limit.
An interesting result of our approach is the description of the transition
Bound States and Superfluidity in Fermion Systems 581
/T\.
1
3.0 - BCS — 1
BEC 1
\
2.5 • 85ft.:." 1
1
2.0
T/e ; /
1.5
1.0
0.5
no
-2.0 -1.5 -1.0 -0.5 0.5 1.0 1.5
Fig. 2. Phase transition region to the superfluid state for a Fermionic system
with separable interaction (17), y = 6.25. The solution of the gap equation
(14) including correlations in the medium is compared with the solution of the
Gorkov equation for vanishing gap (BCS) and the Bose-Einstein condensation of
noninteracting bound states (BEC). Bound states are "blocked out" for densities
above the Mott line.
References
[1] P. Nozieres and S. Schmitt-Rink, J. Low Temp. Phys. 59, 159 (1985);
M. Randeria, this volume.
[2] M. Drechsler and W. Zwerger, Ann. Physik 1, 15 (1992); AJ.
Leggett, in Modern Trends in the Theory of Condensed Matter,
A. Pekalski and J. Przystawa, eds. (Springer, Berlin, 1980). M.
Randeria, J.M. Duan, and L.Y. Shieh, Phys. Rev. Lett. 62, 981
(1989); Phys. Rev. B 41, 327 (1990); see also M. Randeria, this
volume.
[3] H. Haug and S. Schmitt-Rink, Progress in Quantum Electronics 9,
3 (1984); R. Zimmermann, Many-Particle Theory of Highly Excited
Semiconductors (Teubner, Leipzig, 1987).
[4] I.F. Silvera, J. Low Temp. Phys. 89, 287 (1992); D. Vollhardt and
P. Wolfle, The Superfluid Phases of Helium-3 (Taylor and Francis,
London, 1990).
[5] R.K. Su, S.D. Yang and T.T.S. Kuo, Phys. Rev. C 35, 1539 (1987);
J.M.C. Chen, J.W. Clark, E. Krotscheck, and R.A. Smith, Nucl.
Phys. A 451, 509 (1986); T. Aim, G. Ropke, and M. Schmidt, Z.
Phys. A 337, 355 (1990).
[6] W. Weise and U. Vogl, Progr. Part. Nucl. Phys. 27, 195 (1991);
S. Klevanski, Rev. Mod. Phys. 64, 649 (1992).
[7] S.A. Chin and E. Krotschek, Phys. Rev. Lett. 65, 2658 (1990); Phys.
Rev. B 33, 3158 (1986); see also D.M. Ceperley and E.L. Pollock,
Phys. Rev. Lett. 56, 351 (1986); Phys. Rev. B 36, 8343 (1987).
[8] C.E. Campbell and B.E. Clements, in Elementary Excitations in
Quantum Fluids, K. Ohbayashi and M. Watanabe, eds. (Springer,
Berlin, 1989).
Bound States and Superfluidity in Fermion Systems 583
[9] M. Schmidt, G. Ropke, and H. Schulz, Ann. Phys. (N.Y.) 202, 57
(1990).
[10] S. Schmitt-Rink, C M . Varma, and A.E. Ruckenstein, Phys. Rev.
Lett. 63, 445 (1989).
[11] W.H. Dickhoff, Phys. Lett. B 210, 15 (1988); W.H. Voderfecht,
C.C. Gearhart, W.H. Dickhoff, A. Polls and A. Ramos, Phys. Lett.
253, 1 (1991).
[12] G. Ropke and H. Schulz, Nucl. Phys. A 477, 472 (1988).
[13] D.N. Zubarev, Nonequilibrium Statistical Thermodynamics (Plenum,
New York, 1974).
[14] G. Ropke, T. Seifert. H. Stolz, and R. Zimmermann, Phys. Stat. Sol.
(b) 100, 215 (1980).
36
Onset of Superfluidity in Nuclear Matter
A. Hellmich, G. Ropke, A. Schnell, and H. Stein
Fachbereich Physik and MPG-AG "Theoretische Vielteilchenphysik"
Universitat Rostock
PF 999, D 18051 Rostock
Germany
Abstract
Within the Gorkov approach, the onset of superfluidity in nuclear matter
is considered in the spin-singlet (S=0) and triplet (S=l) channel. In the
triplet channel a transition from Bose-Einstein condensation (BEC) of
deuterons at low densities to a BCS state at high densities is obtained. It
is shown that correlations and bound state formation in the medium should
be included to improve the Gorkov approach. The drastic change in the
composition of the system due to the Mott effect is investigated.
1 Introduction
Nuclear matter is an interesting example of strongly coupled fermions
showing a transition to a superfluid state at densities n ^ no = 0.17 fm~3
and temperatures of the order T ~ 1 MeV. We consider symmetric
nuclear matter containing equal ratios of protons and neutrons. The
nucleon-nucleon interaction is taken in the form of a separable poten-
tial [1]
~~ J(*'), (1)
2S+1
where f^(fe) denotes the form factor in the channel a = Lj for angu-
lar momentum L, tffj are coupling constants .
584
Onset of Superfluidity in Nuclear Matter 585
and Cooper pairing (BCS) at higher densities. (See also the article by
Randeria in this volume.) Therefore, we expect a smooth transition from
strong coupling in the BEC regime to weak coupling in the BCS regime.
We apply the formalism given by Nozieres and Schmitt-Rink [2] (see
also [3]) to nuclear matter.
In Section 2 the critical temperature is derived using two alternative
approaches, the equation of motion for thermodynamic Green functions
and the T-matrix approach. For the gap and the critical temperature
model, calculations using the Yamaguchi potential are given in Section
3. The equation of state and the composition of nuclear matter, the Mott
effect and the onset of a superfluid phase are discussed in Section 4. For
convenience we put h, c, kg =1.
{tj V
\ E I I lmWajamal > (2)
where k is the relative and K the total momentum and Ae(k,K) is the
angle averaged quasiparticle energy shift due to the interactions with the
medium. In this approximation, the effects of the medium enter solely
through the phase-space occupation factors /(k, K) and the quasiparticle
energy shift As(k,K).
where
dkk2
A f (6)
The coupling of the 3Si state to the 3D\ state is neglected. The parameters
/?, ks (singlet channel) and Xt (triplet channel) are fitted to the empirical
nucleon-nucleon scattering phase shifts and the deuteron binding energy
El = -2.22 MeV.
With the inverse potential range /?=1.4488 fin"1, empirical values for the
potential parameters 2a are
This simple rank 1 potential has been choosen to illustrate the many-
particle effects in nuclear matter. In order to obtain the effects of
a repulsive core, more realistic nucleon-nucleon interaction should be
considered. Starting from a general BCS theory for pairing in arbitrary
channels [8] this is done in [6] for the Graz-II potential which also
considers the coupling of the 3S\ - 3D\ channel. The fact that different
realistic potentials yield very similar results has been discussed in Refs. [9,
10].
(10)
where C contains the summation over k'. Fig. 1 shows the shape of the gap
which is proportional to the form factor of the potential. At the critical
588 A. Hellmich, G. Ropke, A. Schnell, and H. Stein
A[ MeV]
Fig. 1. Gap function A(/c) for triplet and singlet channel at /xeff = 2 MeV and
T = 0.3 MeV.
+00
1= (11)
where
(12)
where f(k) is the Fermi distribution function which contains the shift
self-consistently.
-10 0 10 20 30 40 50 60 70
Fig. 2. Critical temperature in the 3Si and 1So channel without (dashed) and with
(bold) selfconsistent HF shifts.
) + 2ncon(fl, T)
with
0.8
g 0.6
"free = 4
= 3
r
*scat = -3
J i""*
-3 E)
with a degeneracy factor 3 for the triplet and singlet channels, / is the
Fermi, g is the Bose distribution function, £COnt = K2/4m + 2As(K,k = 0)
is the continuum edge and Sa(E9K9fi,T) are the generalized scattering
phase shifts in the 3S\ and ^o channels. Binding energies Eb are calculated
according to (14). Here the quasiparticle shifts are considered only in a
rigid shift approximation (effective mass m* = m), which represents the
same level of approximation as in Ref. [2] or [3].
As expected from the discussion of the binding energy the Mott effect
already leads, for lower temperatures, to a strong suppression of the
correlated pair density at low densities. Fig. 3 compares this with a
classical Beth-Uhlenbeck calculation (nucleons and deuterons considered
as ideal Boltzmann particles), where the contribution of correlated density
is monotonically increasing with total density. Because the Green's
592 A. Hellmich, G. Ropke, A. Schnell, and H. Stein
5 Conclusion
It has been shown that the solutions of both the Gorkov equation and
the T-matrix equation coincide at the critical temperature for the onset
of superfluidity. In the low-density limit the latter is identical with the
Schrodinger equation for bound states so that we obtain the correct
description in the low-density limit (BEC) as well as in the high-density
Onset of Superfluidity in Nuclear Matter 593
limit (BCS). However, the Gorkov gap equation is not correct in the
entire temperature-density plane because it uses a HF type decoupling
procedure for the higher Green functions, equivalent to the simple ladder
summation with account taken of H F single particle Green functions in
the T-matrix approach. This approximation means all correlations in the
medium are neglected. However, in a system such as nuclear matter, we
have strongly interacting fermions which are correlated. In particular,
they can form bound states (deuterons). Therefore, a H F type approach
is not always justified.
Thus, we use a generalized Beth-Uhlenbeck formula as an equation
of state for the normal phase to calculate the composition of the system.
It has been found that bound states are of major importance in the
region of low temperatures and low densities. With increasing density,
bound states disappear due to the Pauli blocking (Mott effect) so that
at high densities compared with the Mott density, the approximation of
uncorrelated quasiparticles is possible.
To consider bound states in the medium, the Gorkov equation needs
to be improved. A possible approach with Green functions could be a
cluster Hartree-Fock calculation as performed in [14]. The correlated
medium enters via effective phase-space occupation and cluster-Hartree-
Fock shifts. In the region considered, a liquid-gas phase transition
occurs, as shown in [14], which is important for the investigation of
phase instabilities. For more realistic calculations of nuclear matter,
the Yamaguchi potential has to be replaced by a more sophisticated
nucleon-nucleon interaction. Further investigations should also include
larger clusters such as tritons and a-particles.
The authors thank T. Aim for his valuable suggestions and support of
this work.
References
[1] J. Haidenbauer and W. Plessas, Phys. Rev. C 30, 1822 (1984).
[2] P. Nozieres and S. Schmitt-Rink, J. Low Temp. Phys. 59, 195 (1985).
[3] M. Schmidt, G. Ropke, and H. Schulz, Ann. Phys. 202, 57 (1990).
[4] R.K. Su, S.D. Yang, and T.T.S. Kuo, Phys. Rev. C 35, 1539 (1987).
[5] T. Aim, GSI-93-07 Report (dissertation), (1993).
[6] T. Aim, B. Friman, G. Ropke, and H. Schulz, Nucl. Phys. A 551, 45
(1993).
[7] Y. Yamaguchi, Phys. Rev. 95, 1628 (1954).
[8] R. Tamagaki, Prog. Theor. Phys. 44, 905 (1970).
594 A. Hellmich, G. Ropke, A. Schnell, and H. Stein
[9] M. Baldo, J. Cugnon, A. Lejeune, and U. Lombardo, Nucl. Phys. A
515, 409 (1990).
[10] M. Baldo, I. Bombaci, and U. Lombardo, Phys. Lett. B 283, 8
(1992).
[11] M. Randeria, J.M. Duan, and L.Y. Shieh, Phys. Rev. Lett. 62, 981
(1989).
[12] M. Randeria, this volume.
[13] R.F. Bishop, H.B. Ghassib, and M.R. Strayer, Phys. Rev. A 13, 1570
(1976).
[14] G. Ropke, M. Schmidt, L. Miinchow, and H. Schulz, Nucl. Phys. A
379, 536 (1982) and A 399, 587 (1983).
Appendix. BEC 93 Participant List
International Workshop on Bose-Einstein Condensation
(Levico Terme, 31 May - 4 June, 1993)
595
596 Appendix. BEC 93 Participant List
Franck Laloe, Laboratoire de Physique de TENS, France
Andrea Lastri, Universita di Trento, Italy
Anthony Leggett, University of Illinois at Urbana-Champaign, USA
R. Leonardi, Universita di Trento, Italy
Jia Ling Lin, University of Illinois at Urbana-Champaign, USA
Bennett Link, Los Alamos National Laboratory USA
Akira Matsubara, Kyoto University, Japan
S.A. Moskalenko, Academy of Sciences of Moldova, Moldova
A. Mysyrowicz, Ecole Polytechnique, France
Nobukata Nagasawa, University of Tokyo, Japan
Fabio Pistolesi, Scuola Normale Superiore, Pisa, Italy
Ludovic Pricavpenko, Universite Paris-Sud, Orsay, France
Mohit Randeria, Argonne National Laboratory, USA
Julius Ranninger, CNRS-CRTBT, Grenoble, France
John D. Reppy, Cornell University, USA
Meritt Reynolds, University of Amsterdam, The Netherlands
Mannque Rho, Saclay, France
Gerd Ropke, Universitat Rostock, Germany
G.V. Shlyapnikov, Kurchatov Institute, Russia
L.D. Shvartsman, The Hebrew University of Jerusalem, Israel
Yu.M. Shvera, Academy of Sciences of Moldova, Moldova
Isaac Silvera, Harvard University, USA
David Snoke, The Aerospace Corporation, Los Angeles, USA
Paul E. Sokol, Penn. State University, USA
Holger Stein, Universitat Rostock, Germany
H.T.C. Stoof, Eindhoven University of Technology, The Netherlands
Sandro Stringari, Universita di Trento, Italy
Eric Svensson, AECL Research, Canada
E. Tiesinga, Eindhoven University of Technology, The Netherlands
Sergei Tikhodeev, General Physics Institute, Russia
Jacques Treiner, Universite Paris-Sud, Orsay, France
D. van der Marel, University of Groningen, The Netherlands
Silvio A. Vitiello, Universita degli Studi di Milano, Italy
Jook T.M. Walraven, University of Amsterdam, The Netherlands
Tilo Wettig, S.U.N.Y. at Stony Brook, USA
Jim Wolfe, University of Illinois at Urbana-Champaign, USA
Index
597
598 Index
biexciton-biexciton interaction, 335-336, charged bosons, 377, 380-381, 395, 411,
504 532, 544-547
binding energy, 333 chiral symmetry, 3, 439
condensation of, 265, 337, 340-346, coherence length, 18, 357, 371, 395, 546
487^96, 505, 507, 510 coherent pairing, 507, 510-511
creation, 333-334, 345 coherent potential approximation, 367
Lamb shift, 497-504 collisions, 204, 208, 214, 215, 228
luminescence, 337-344, 488, 511 atoms, 103, 190, 465^70
momentum distribution, 338 biexcitons, 249
s-wave scattering length, 249, 270, 494 excitons, 295, 296, 300, 308, 312, 318, 513
4
single-photon absorption, 333 He, 66
two-photon absorption, 265, 333-334, positronium, 561-564
339,489,496,508,511 spin-polarized hydrogen, 134-136,
two-photon recombination, 342 139-140, 143, 145, 147, 148, 161,
biholes, 532-539 227, 228
Bil 3 , 533 composite bosons, 1, 4, 24-25, 247, 253,
binary collision approximation, 101 283, 355, 356, 358, 371, 387, 423^25,
bipolarons, 22, 358, 395^14, 541-547 465, 532
condensation of, 395, 397, 400, 407-413, condensate
542-547 amplitude fluctuations, 20, 205, 210, 243
bismuth films, 28 in momentum distribution, 55, 58, 61, 68,
bismuth oxide superconductors, 397 100, 206, 220, 234, 294, 302
black holes, 444-450 phase coherence, 21, 22, 25, 32-35
Bogoliubov approximation, 21, 44, 101, phase fluctuations, 8, 34, 35, 46, 203, 205,
335, 379 217, 220, 243
Bogoliubov dispersion law, 92, 209, 220, wavefunction, 32, 33, 205, 234, 456
243, 374, 544 Cooper pairs, 5, 249, 258, 334, 355-358,
Bogoliubov inequality, 87, 89, 91 370, 373, 375, 387, 388, 394, 395, 421,
Bogoliubov sum rule, 87 453, 455^58, 507, 532, 572, 585
Bogoliubov transformation, 40, 466, 510, in disordered alloys, 27
correlation length, 163, 205
576
CP violation, 226
Bohr solution of hydrogen atom, 282,
critical exponents, 4, 46, 48, 79
331-332
Cu 2 O, 6, 21, 253, 281-326, 346-353, 439,
Boltzmann equation, 202, 204, 206-208,
442, 519-523, 525, 526
227, 228, 236, 239, 277, 544, 575
band structure constants, 283
Bose gas, see weakly interacting Bose gas,
exciton lifetime, 286, 288, 346-347
hard sphere Bose gas
excitonic relaxation processes, 300
Bose glass, 26, 27, 43, 46
optical phonons, 286
Bose liquid, 18, 22, 58, 160, 265, 358,
CuCl, 7, 336-346, 487-495, 504, 508, 511
551-557 cuprate superconductors, 358, 382, 397,
condensate fraction, 57 414, 542, 547
driven, 264 CuO 2 planes, 397^01, 413, 541
hydrodynamic equations, 221
phonons, 89, 552, 554, 555
DeBroglie wavelength, 53, 105, 137, 145,
Bose narrowing, 6, 293
190, 198, 235, 293, 466
Bose-Einstein distribution, 204, 292, 293,
dilute Bose gas, see weakly interacting
302, 305, 314, 340, 347, 488, 524 Bose gas
bosonization, 9, 418^36, 570, 571 Doppler cooling, 176-177, 179
broken gauge symmetry, 1-3, 7, 21, 32-35,
52, 93, 234, 236, 241, 368, 372, 376,
electron-hole droplets, see electron-hole
377, 387, 439, 452-462, 567
liquid
BSCO, 397, 412, 547
electron-hole exchange, 262, 286, 335, 346
electron-hole insulator, 254
cesium, see atomic cesium electron-hole liquid, 247, 249-252, 254,
charge density wave, 252, 361 287, 335, 358, 528
charge transfer gap, 397-399 electron-hole plasma, 249
Index 599
electron-phonon coupling, 394-395 Fermi-Dirac distribution, 254, 261
electronic phase transition, 252 of localized bosons, 8, 546
Eliashberg theory, 395 four-wave mixing, 265, 273-274, 496
entropy loss, 312, 442 phase conjugation, see optical phase
exchange effects, 17, 103-110, 119-122, 425, conjugation
435 fractional quantum Hall effect, 90
excitonic crystal, 249 fullerenes, 397, 414, 541
excitonic gas, 247-255, 263, 281-284, 288,
295-325, 438 GaAs, 500, 502, 508, 534, 535, 537
excitonic insulator, 249, 250, 257 Galitsky method, 103
excitonic molecule, see biexcitons GaP, 534
excitons, 7, 10, 18, 21, 24, 25, 82, 100, 202, GaSb, 534
246-277, 281-326, 330-353, 355, 426, Ge, 250, 253, 293, 323, 525, 526, 528, 534,
439, 442, 487, 519-530 538
Auger process, 296, 300, 308, 311, 315, Gell-Mann-Levy model, 442
325 Ginzburg-Landau theory, 33-35, 442, 527
binding energy (excitonic Rydberg), 248, time-dependent, 238, 368 -372
282-283, 285, 331 Goldstone bosons, 568
Bohr radius, 248, 282, 331 gravitational atomic cavity, 194-197
condensation of, 6, 39, 100, 203, 254, Gross-Pitaevsky equation, see nonlinear
302-305, 307, 312, 316-320, 332, Schrodinger equation
334-353, 357, 377, 507, 510, 520-523 gun, smoking, see BEC, direct evidence for
creation, 253, 265, 284, 295, 296, 332 hard core Bose gas, see hard sphere Bose
diffusion, 289-291, 296, 320, 332, 349, gas
520, 528 hard sphere Bose gas, 23, 38^42, 44, 56, 64,
dumbbell, 8 100, 102, 117-118, 284, 297, 361, 379,
exciton-exciton interaction, 263-264, 553, 565
269, 296, 297, 312, 320, 326, 335, Hartree-Fock approximation, 17-19, 103,
501, 507-510, 513-516, 526 121, 237, 577, 580, 593
exciton-phonon interaction, 288-291, healing length, 148, 149
298-300, 305, 310, 320, 336, 352, 3
He, 18, 485
526, 529 superfluid, 5, 355, 357, 453, 455, 457, 574
in a stress field, 291, 313-325 4
He, 6, 35-38, 42, 51-82, 92, 160, 162, 168,
in two dimensions, 8, 270 248, 254, 286, 355, 478, 550, 551
luminescence, 284^325, 347-349 gas, 250
momentum distribution, 286, 298, in a porous medium, 10, 27, 43, 46
301-320 in two dimensions, 8
ortho, 21, 286-325, 332, 346-349, 507 lambda transition, 42
para, 21, 286-325, 332, 346, 349, 507, momentum distribution, 4, 53, 58-82,
508, 519, 525 286, 551
phonon-assisted spin relaxation, 288, correlated basis function calculation,
300, 302, 314 60
photovoltaic detection, 349-352, 519 Green's function Monte Carlo
polarized, 508, 510 calculation, 60-62, 64, 74, 76, 77,
"quantum saturation", 21, 223, 307-312, 80
314^315, 325, 349 path integral Monte Carlo calculation,
recombination, 9, 252, 253, 284-289, 298 60, 62, 70, 73, 77, 79, 81
s-wave scattering length, 297, 312 shadow wave function calculation, 60
single-photon absorption, 331-332 variational calculation, 60, 61, 64
superfluidity of, 7, 39, 203, 252-253, 305, phase diagram, 56
311, 334-335, 349-353, 520-523, solid, 23, 56, 80
528, 529 superfluid, 1, 9, 135, 142, 161, 221, 229,
two-body spin relaxation, 21, 300, 308 355, 459
virtual, 262 condensate fraction, 62, 64, 68-81, 101
maxons, 96
Fermi gas, see attractive Fermi gas phonons, 35, 36, 60, 61, 63, 64, 96
Fermi liquid, 247, 249, 259, 358-367, 404 rotons, 63, 96
600 Index
Higgs boson, 3 magnetoexcitons, 8
Hohenberg-Mermin-Wagner theorem, maser, 132
87-90 Maxwell-Boltzmann distribution, 54, 288,
holons, 35 289, 298, 299, 302, 314, 338, 347, 488
Hubbard model, 409, 413 mean-field approximation, 23, 34, 100, 101,
attractive, 360-361, 367, 373, 377, 379, 121, 259, 260, 269, 271, 419, 575
381, 387 Mott transition, 26, 27, 575, 581, 590, 592
repulsive, 358, 381
hydrogen, see spin-polarized hydrogen neutrinos, 441, 447
molecules, 142, 332-333, 335 neutron scattering, 4, 53, 62-81, 550-557
under high pressure, 359 final-state effects, 66-68, 71, 73, 74
hyper-Raman scattering, 511 impulse approximation, 63, 64, 66,
550-552, 555, 556
InAs, 534
neutron stars, 439, 444
InP, 534
nonequilibrium Green's functions
InSb, 534
interacting boson model, 418, 426, 427 electron-hole, 257-262
Ising model, 23, 567 exciton, 264-269
nonequilibrium phase transition, 255, 514
Josephson effect, 21 nonlinear optics, 255, 257, 262, 263, 265,
Josephson junction, 454, 455, 458 277
nonlinear Schrodinger equation, 34, 103,
kaons, 438-450 205, 207, 242, 371, 545
kaon-nucleon interaction, 439-440 nuclear ferromagnetism, 19
kaonic atoms, 439 nuclear matter, 418, 574, 582, 584-593
Kosterlitz-Thouless transition, 4, 8, 18, 35, equation of state, 359, 585
38, 163, 203, 368 nuclear spin waves, 104, 132
of 4 He, 8 nucleons, 3, 418-436
of excitons, 8 nucleon-nucleon interaction, 584, 587
of spin-polarized hydrogen, 8, 168-170, nucleosynthesis, 448-449
478^85
off-diagonal long-range order, 1, 4, 8, 33,
Levy flights, 194 57, 59, 206, 220-223, 356, 368, 387
lambda transition, see 4 He, lambda one-body density matrix, 32, 57, 60, 94,
transition 100, 104, 109, 122
Landau-Ginzburg theory, see Onsager-Feynman quantization, 33
Ginzburg-Landau theory optical lattices, 180, 186-187
laser, 4, 9, 255
optical molasses, 176, 179, 185-188, 194
and BEC, see BEC to laser crossover
optical phase conjugation, 344-346,
excitonic, 255-257
488-495
neutrino, 9
laser cooling of atoms, 100, 166, 173-198,
202, 466 Pauli exclusion, 253, 254, 292, 379, 387,
Doppler cooling, 154 425, 579-582, 590
recoil limit, 190 phase locking, see condensate, phase
Sisyphus effect, 179, 182-186, 189 coherence
velocity-selective trapping, 190-194 phase transition
laser generation of excitons, see excitons, first order, 32, 41, 42
creation second order, 34, 41, 79, 229-232, 443
linear response theory, 36, 38 phase-space filling, 262
lithium, 227 phonon wind, 528-530
local density approximation, 152 phonons, 220, 222
long-range phase coherence, see in Bose liquid, see Bose liquid, phonons
off-diagonal long-range order in crystal lattice, see electron-phonon
LSCO, 397 coupling, exciton-phonon
interaction
magneto-optical trap, 166, 177-179, in superfluid helium, see 4 He, superfluid,
188-190 phonons
Index 601
in weakly interacting Bose gas, see self-consistent field approximation, 266,
weakly interacting Bose gas, 515, 516
phonons self-phase modulation, 265
plasmons, 377, 381, 388, 559 shell model, 419, 426
polariton bottleneck, 256 Si, 250, 253, 293, 323
polaritons, 255-262, 268-269, 336, 338, 342, SiGe, 538
497-504, 513-516 Silvera-Reynolds number, 525
condensation of, 9, 257, 513-516 single-particle density matrix, see one-body
dispersion law, 255, 339 density matrix
polarons, 395, 398 smoking gun, see BEC, direct evidence for
positronium, 7, 283, 285, 287, 300, 331 solitons, 353
condensation of, 558-564 spin density wave, 252
molecule, 333, 559, 564 spin gaps, 358, 382-387, 414, 542
momentum distribution, 563 spin waves, 565-568
s-wave scattering length, 562 spin-polarized hydrogen, 7, 18, 19, 100,
superfluidity of, 564 131-157, 160-171, 202, 203, 226-228,
two-body spin relaxation, 562 241, 295, 442, 453, 525, 574
two-photon recombination, 563 compression, 136
"precondensate", 211 condensation of, 39, 100, 131-156,
pseudogap, 407, 408, 410, 413, 414, 543 160-161, 163-167, 227, 228,
pseudopotential method, 101 231-232, 236, 241
evaporative cooling, 138-147, 154, 163,
QCD theory, 439 165, 227, 236, 475
quantum heterostructures, 8, 533 hyperfine splitting, 133, 161, 227
quantum lattice model, 565 in two dimensions, 162-163, 168-170,
quantum Monte Carlo method, 381, 382 478^85
quantum tunneling, 454 magnetic trapping of, 134-157, 161, 227,
quantum well, 8, 10, 270, 500, 508, 509, 533 472-475
quantum wire, 502 microwave trapping of, 163-167
quasicondensate, 8, 163, 168, 203, 205, relaxation explosion, 151, 164, 165,
208-211, 218, 220, 223, 243, 320, 527 472-475
quasiequilibrium, 213, 253, 267, 295, 300, s-wave scattering length, 145, 148-149,
336 227
"quasipairs", 126 superfluidity of, 39, 132, 160
quasiparticles, 35, 36, 38, 40, 41, 119, 122, three-body recombination, 132, 134, 137,
242, 378, 591 161, 165, 169, 210-211, 472, 479-484
two-body spin relaxation, 135, 140, 143,
Rabi oscillations, 257, 275, 277 147, 162, 165, 300
radiation pressure, 175 spontaneous symmetry breaking, see
random phase approximation, 204-206, broken gauge symmetry
377, 378, 380, 381, 383 stimulated emission, 6
stimulated scattering, 6, 273, 292, 344, 352
s-d boson model, see interacting boson strangeness condensation, 443
model strongly interacting Bose gas, 160
s-wave scattering length, 39, 126, 211, 237, superconductivity, 3, 5, 24, 25, 249, 355,
360, 362, 371 356, 361, 374, 394, 403, 421, 426, 453,
atomic cesium, see atomic cesium, 532, 574
s-wave scattering length collective excitations, 377-381
biexcitons see - biexcitons, s-wave high T c , 5, 35, 357-358, 382, 386-387,
scattering length, 249 394-414, 462, 541-547, 569
excitons, see excitons, s-wave scattering and bipolarons, see bipolarons,
length condensation of
positronium, see positronium, s-wave in disordered system, 10, 358
scattering length Schafroth, 395, 396
spin-polarized hydrogen, see type II, 10
spin-polarized hydrogen, s-wave superfluid density, see two-fluid model,
scattering length superfluid density
602 Index
superfluid gas, 128 biexcitons, see biexcitons, two-photon
superfluidity, 3-5, 8, 21, 23-25, 27, 33, recombination
35-38, 4 2 ^ 8 , 51, 59, 87, 101, 127, 160, excitons, 285
163, 206, 207, 217, 220, 223, 355, 356, positronium, 285, 559
387, 407, 453, 455-462, 574-582,
584-593 ultrafast optics, 298
supernova 1987A, 440, 449, 450 "universal phase standard", 455, 459
supersolids, 23 Ursell operators, 102-103, 110-128
symmetry breaking, see broken gauge
symmetry vortices, 21, 34, 35, 206, 217, 219, 220, 222,
223
thermodynamic limit, 9, 16, 117-118, 228, weakly interacting Bose gas, 5-7, 82, 91-93,
235, 243, 452, 457, 526 95, 100-128, 148-149, 160, 163, 168,
topological long-range order, 206, 217-220, 203, 204, 226, 284, 291-294, 325, 336,
222 371, 372, 380, 438, 524, 526
two-body density matrix, 94-97, 101 blue shift, 320, 494
two-body spin relaxation equation of state, 110-118
of excitons, see excitons, two-body spin far from equilibrium, 203-224, 227-243,
relaxation 514
of hydrogen, see spin-polarized phonons, 206, 218, 220-223, 388
hydrogen, two-body spin relaxation speed of sound, 149, 218, 221, 379
of positronium, see positronium, weakly interacting Fermi gas, 357, 360
two-body spin relaxation Wigner crystal, 380
two-fluid model, 4, 36-38, 221-222
superfluid density, 41, 43, 45-47, 81, 368 YBCO, 382, 386, 387, 397, 399^01, 412,
two-photon absorption 542-545
by biexcitons, see biexcitons, two-photon
absorption ZnP 2 , 533
by hydrogen, 154-157 ZnSe, 508
two-photon recombination ZnTe, 534