-
Notifications
You must be signed in to change notification settings - Fork 1
/
thesis_berceanu.txt
6023 lines (4721 loc) · 241 KB
/
thesis_berceanu.txt
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
Superfluidity, the ability of a fluid to flow without apparent
viscosity, is one of the most striking consequences of collective
quantum coherence, with manifestations ranging from metastability of
supercurrents in multiply connected geometries to the appearance of
quantized vortices, or the existence of a critical velocity for
frictionless flow when scattering against a defect. While
traditionally investigated in equilibrium systems, like liquid
He and ultracold atomic gases, experimental advances in
nonlinear optics, in particular regarding microcavity
exciton-polaritons, paved the way for studying superfluid-related
phenomena in a driven-dissipative framework.
Equally exciting is the possibility of realising topological phases of
matter, such as the integer or fractional quantum Hall states, outside
of traditional electronic systems. Photonics experiments in
driven-dissipative resonator arrays, in particular, offer a high
degree of controllability and tunability, as well as unprecedented
experimental access to the eigenstates and energy spectrum.
This thesis reports on hydrodynamic effects, as well as topological
properties, of driven-dissipative systems. In particular, we analyze
the superfluid-like behaviour of microcavity exciton-polaritons, as
well as the momentum-space topology of coupled resonator arrays.
Microcavity exciton-polaritons are quasiparticles resulting from the
mixing of excitons (bound electron-hole pairs) and photons confined
inside semiconductor microcavities. While polariton fluids have been
shown to display collective coherence, the connection between the
various manifestations of superfluid behaviour is more involved
compared to equilibrium systems. In this manuscipt, we consider both
the case of a single-fluid pump-only configuration, as well as the
three-fluid optical parametric oscillator regime that results from
parametric scattering of the pump to the signal and idler states. In
both cases, we look at the response of the moving polaritons
scattering against a weak static defect present in the microcavity.
For the single fluid, we evaluate analytically the drag exerted by the
fluid on the defect. For low fluid velocities, the pump frequency
classifies the collective excitation spectra in three different
categories: linear, diffusive-like and gapped. We show that both the
linear and diffusive-like cases share a qualitatively similar
crossover of the drag from the subsonic to the supersonic regime as a
function of the fluid velocity, with a critical velocity given by the
speed of sound found for the linear regime. In contrast, for gapped
spectra, we find that the critical velocity exceeds the speed of
sound. In all cases, we show that the residual drag in the subcritical
regime is caused by the nonequilibrium nature of the system. Also,
well below the critical velocity, the drag varies linearly with the
polariton lifetime, in agreement with previous numerical studies.
The optical parametric oscillator regime presents an additional
challenge, as one is dealing with three coupled fluids. The
spontaneous macroscopic coherence following the phase locking of the
signal and idler fluids has been already shown to be responsible for
their simultaneous quantized flow metastability. We find that the
modulations generated by the defect in each fluid are not only
determined by its associated scattering ring in momentum space, but
each component displays additional rings because of the cross-talk
with the other components imposed by nonlinear and parametric
processes. We single out three factors determining which one of these
rings has the biggest influence on each fluid response: the coupling
strength between the three fluids, the resonance of the ring with the
polariton dispersion, and the values of each fluid group velocity and
lifetime together establishing how far each modulation can propagate
from the defect. For the typical conditions of parametric scattering,
the pump is in the supercritical regime, so the signal and idler will
show the modulations of the pump, meaning none of the three states
manifests superfluid behaviour. However, the signal appears to flow
without friction in the experimental study, because the three factors
mentioned above conspire to reduce the amplitude of its modulations
below currently detectable levels.
Driven-dissipative systems can show interesting phenomena also without
interactions, stemming from the nontrivial topology of their energy
bands. In the final part of this thesis, we present a realistic
proposal for an optical experiment using state-of-the-art coupled
resonator arrays. We study theoretically the driven-dissipative
Harper-Hofstadter model on a square lattice in the presence of a weak
harmonic trap. Without pumping and losses, the eigenstates of this
system can be understood, under certain approximations, as
momentum-space toroidal Landau levels, where the Berry curvature, a
geometrical property of an energy band, acts like a momentum-space
magnetic field. We show how key features of these eigenstates can be
observed in the steady-state of the driven-dissipative system under a
monochromatic coherent drive. We also show that momentum-space Landau
levels would have clear signatures in spectroscopic measurements in
such experiments, and we discuss the insights gained in this way into
geometrical energy bands and particles in magnetic fields.
List of Publications
List of Publications
The following articles have been published in the context of this thesis:
Onset and dynamics of vortex-antivortex pairs in polariton optical parametric oscillator superfluids,
G. Tosi, F. M. Marchetti, D. Sanvitto, C. Anton, M. H. Szymanska, A. C. Berceanu, C. Tejedor, L. Marrucci, A. Lemaitre, J. Bloch, and L. Vina,
Phys. Rev. Lett. 107, 036401 (2011).
Drag in a resonantly driven polariton fluid,
A. C. Berceanu, E. Cancellieri, and F. M. Marchetti,
J. Phys.: Condens. Matter 24, 235802 (2012) [Chapter ].
Multicomponent polariton superfluidity in the optical parametric oscillator regime,
A. C. Berceanu, L. Dominici, I. Carusotto, D. Ballarini,
E. Cancellieri, G. Gigli, M. H. Szymanska, D. Sanvitto,
and F. M. Marchetti,
Phys. Rev. B 92, 035307 (2015) [Chapter ].
Momentum-space Landau levels in driven-dissipative cavity arrays,
A. C. Berceanu, H. M. Price, T. Ozawa, and I. Carusotto,
Phys. Rev. A 93, 013827 (2016) [Chapter ].
Conference proceedings:
Momentum-space Landau levels in arrays of coupled ring resonators,
H. M. Price, A. C. Berceanu, T. Ozawa, and I. Carusotto,
Proc. SPIE 9762, 97620W (2016).
Preface
Preface
How birds fly together
Systems far from thermal equilibrium frequently show novel features
when compared to their equilibrium counterparts. As an everyday
example, consider a group of birds displaying long-range ordered
behaviour, manifested by forming a flock under certain
conditions. This behaviour can be modeled by introducing a time step
rule, such that each individual bird in a group determines its next
direction on each time step by averaging the directions of its
neighbours and adding some random noise on top of
that . It can be shown that, in the limit of the
velocity going to zero, the model reduces to the XY model in two
dimensions, where the spin is represented by the bird velocity. Since
the 2D XY model does not spontaneously break the symmetry at any
finite temperature (as justified by the Mermin-Wagner theorem), one
can show that the appearance of the long-range ordered phase is a
direct consequence of nonequilibrium aspects of the model. In a
nutshell, the neighbours of one particular bird will be different at
different times, depending on the velocity field. This gives rise to a
time-dependent variable-ranged interaction, which can stabilize the
ordered phase.
Driven-dissipative systems
Non-equilibrium driven-dissipative photonic systems, such as
polaritons in semiconductor microcavities or arrays of coupled optical
resonators, have recently attracted a lot of interest due to the
possibility of observing quantum phenomena which normally require very
low temperatures and/or intense magnetic fields and are traditionally
restricted to the domain of solid-state systems or ultracold atomic
gases. Besides being highly tunable, these optical systems facilitate
direct experimental access to observables such as the wavefunction or
energy spectrum, all at room temperature. In particular, microcavity
polaritons have allowed the observation of collective hydrodynamic
phenomena, ranging from frictionless flow around a small defect to the
formation of quantized vortices and dark solitons at the surface of
large impenetrable obstacles , while
ring-resonator arrays coupled to artificial magnetic fields have
recently allowed engineering topological edge states robust to
disorder .
Polaritons
Microcavity polaritons are quasiparticles resulting from the strong
coupling of cavity photons and quantum well excitons
, and have the prerogative of being both easy to
manipulate, via an external laser, and detect, via the light escaping
from the cavity . The finite polariton lifetime
establishes the system as intrinsically out of equilibrium: an
external pump is needed to continuously replenish the cavity of
polaritons, that quickly, on a scale of tens of picoseconds,
escape. The pumping can be done resonantly, close to the polariton
energy dipersion, or non-resonantly.
Incoherent pumping: polariton BEC
The landmark observation of polariton
condensation in 2006, as shown in Fig. , was achieved
using a non-resonant experimental setup. In the experiments, the
system was incoherently excited by a laser beam tuned at a very high
energy. Relaxation of the excess energy
lead to a population of the cavity polariton states
and, above a certain laser power threshold, to Bose-Einstein
condensation into the lowest polariton state.
Experimental observation of polariton Bose-Einstein
condensation. A sharp and intense peak corresponding to the lowest
momentum state is formed in the center of the far-field emission [top
panels], with increasing pump power (left to right). The
corresponding energy-resolved emission [bottom panels] show that
above the condensation threshold, the emission comes almost entirely
from the lowest energy state situated at the bottom of the polariton
dispersion. From Ref. .
Coherent pumping: OPO
Due to their energy dispersion and strong nonlinearity inherited from
the excitonic component, polaritons resonantly injected by the
external laser into the pump state with a suitable wavevector and
energy can undergo coherent stimulated scattering into two conjugate
states , called the signal and
the idler, in a process known as optical parametric oscillator
(OPO). Since their first realisation
, the
interest in microcavity optical parametric phenomena has involved
several fields of fundamental and applicative
research
.
Top panels: Simulated (right panel) and observed (left panel)
edge state propagation in a ring-resonator array. The light enters
from one corner and exits from the other, as signaled by the
arrows. Adapted from Ref. . Bottom
panels: Simulated edge state propagation in a polariton system.
Intensities of the exciton (left panel) and photon (right panel)
fields obtained when pumping a lattice at the center of its lower
edge. From Ref. .
Superfluidity
The superfluid properties of a resonantly pumped polariton quantum
fluid, both in the pump-only configuration (without parametric
scattering) as well as the OPO regime, have been actively investigated
experimentally, as well as theoretically . In
particular, a supression of scattering in the pump-only case was
observed below a critical velocity, similar to what
has been predicted by the Landau criterion for equilibrium superfluid
condensates.
While in equilibrium condensates different aspects of superfluidity
are typically closely related , this is no longer
true in a non-equilibrium context. Independent of the pumping scheme,
the driving and the polariton finite lifetime force one to reconsider
the meaning of superfluid behaviour, when the spectrum of collective
excitations is complex rather than real, raising question about the
applicability of a Landau criterion .
An additional complexity characterises the OPO regime, namely, the
simultaneous presence of three oscillation frequencies and momenta for
pump, signal and idler correspondingly increases up to 12 (or 6,
depending on the approximation used to describe the system) the number
of collective excitation branches . Note that from
the experimental point of view, pioneering
experiments have observed a ballistic nonspreading
propagation of signal/idler polariton wavepackets in a triggered-OPO
configuration, as well as demonstrating the existence and
metastability of vortex configurations in the
signal and idler.
Topology in polaritons
Very recently, the polariton community started to explore topological
effects in polariton lattices
. The idea is to break time reversal symmetry by
the application of a strong external magnetic field, giving rise to
energy bands with nontrivial topology. In particular,
Ref. proposes an exciton-photon coupling with
a winding phase in momentum space, giving rise to polaritonic bands
with chiral edge modes that allow unidirectional propagation,
protected against backscattering. The bottom panels of
Fig. show the exciton and photon fields
traveling together as a unique polaritonic counter-clockwise chiral
edge mode. In this context, dissipation in the form of photon losses
provides a coupling to the outside continuum of modes and, as such,
can be used to detect or inject edge modes.
It would be interesting to study the feasability of coupled
micropillars for this type of physics, in view of the recent
experimental findings of a linear graphene-like
dispersion as well as edge
states in honeycomb micropillar lattices.
On the other hand, topologically protected edge states have already
been experimentaly observed in the case of
silicon-based optical resonator arrays (see top panels of
Fig. ), due to recent advances in creating
synthetic gauge fields, which have opened new horizons for simulating
topological phases of matter also with neutral particles, such as
photons (or ultracold
atoms
). Rather than simply replicating previous
measurements, experiments with synthetic gauge fields allow for
unprecedented access to properties such as the eigenstates or
eigenspectrum, while the tunability and controllability of these
experiments offer the prospect of simulating novel physics.
Contents of this thesis
Thesis layout
This manuscript is split in two parts: part I is a detailed review of
the basic concepts, including both the theoretical formalism, as well
as the relevant experiments, necessary for understanding part II,
which presents the three main works published as part of my PhD
(Chapters , and ). The two
systems chosen for ilustrating the basic physical concepts are
ultracold atomic gases (Chapter ) and microcavity
exciton-polaritons (Chapter ). We now give a brief
description of the content of each of the chapters.
In Chapter , we describe the phenomenon of
Bose-Einstein condensation, and introduce the Gross-Pitaevskii
equation which is extended in Chapter to the case
of polariton condensates. We then review the linear response of a
moving atomic condensate to a weak stationary defect, leading to
discussion on superfluidity. This scattering problem is the
equilibrium conterpart of the problems studied in
Chapters and in the context of polaritons
in the resonantly pumped and optical parametric oscillator regimes,
respectively. Finally, we introduce the Harper-Hofstadter model, which
is the backbone of Chapter .
In Chapter , we introduce microcavity
exciton-polaritons, and their theoretical description in terms of a
driven-dissipative Gross-Pitaevskii equation. We discuss both the
resonantly pumped system which is used in Chapter , as
well as the optical parametric oscillator of Chapter ,
and we make the connection to the superfluid-related phenomena
previously introduced in Chapter .
In Chapter , we study the scattering of a resonantly
pumped polariton condensate against a static defect of the
microcavity, for the simplified case of small fluid velocities, when
the polariton dispersion can be considered quadratic. After discussing
the effects of dissipation on the Bogoliubov excitation spectra, we
find that the finite polariton lifetime also affects the drag exerted
by the condensate on the defect. In particular, we show that there is
a nonzero drag force, entirely due to the out-of-equilibrium nature of
the system, even in the "superfluid regime". Finally, we
characterise the behaviour of the drag as a function of the condensate
velocity and polariton lifetime.
In Chapter , we present a theoretical and experimental
study of the same scattering problem, now in the context of an optical
parametric oscillator consisting of three coupled condensates: the
pump, the signal and the idler. Apart from using the linear response
analysis first introduced in Chapter and then
extended to the case of a pump-only condensate in
Chapter , we also numerically solve the full
driven-dissipative Gross-Pitaevskii equation. We show that, while the
modulation patterns are present in all three condensates, their
relative amplitudes depend on various factors. In particular, for the
typical experimental conditions that favour parametric scattering, the
signal has undetectable modulations, while the pump and idler show the
same response.
In Chapter , we add a harmonic trap to the
Harper-Hofstadter lattice model of Chapter . We
discuss how the eigenstates of the new Hamiltonian can be seen as
momentum-space Landau levels, by making use of the nontrivial topology
of the Harper-Hofstadter energy bands. We then extend the model to
include driving and dissipation, and finally present a proposal for
the physical implementation of the model in state-of-the-art
driven-dissipative photonic systems.
The source code belonging to Chapters , and is available online .
Condensed-matter systems
Ultracold atomic gases
Laser cooling allowed
achieving temperatures in the micro-Kelvin regime, and eventually led
to the realization of optical lattices . It
also paved the way for more powerful cooling techniques, such as
evaporative cooling, which made possible the Bose-Einstein
condensation of dilute atomic
gases .
Ultracold gases have been at the forefront of simulating quantum
phenomena with analogs throughout physics, from nonlinear optics to
condensed matter systems . In particular,
their link to quantum simulation of condensed matter phenomena becomes
obvious when adding optical lattice
potentials , combined with synthetic magnetic
fields .
In this Chapter, we review the physics of ultracold atomic gases, with
an emphasis on their link to hydrodynamic effects such as
superfluidity, as well as their connection to traditional solid state
lattice models such as the celebrated Harper-Hofstadter model which
originally describes the single-particle physics of band electrons in
intense magnetic fields.
Chapter organization
This Chapter is organized as follows: in Section we
present a short history of Bose-Einstein condensation, describe its
main features and state its formal definition in terms of the
Penrose-Onsager criterion, leading to the weakly-interacting Bose gas
paradigm and the Gross-Pitaevskii equation (Sec. ), an
essential theoretical tool for the mean-field description of atomic
condensates. In Section we introduce the
linear response formalism, which proves useful for interpreting
experiments where a weak perturbation is applied to the condensate. We
employ this formalism in order to study a scattering problem
concerning the flow of a BEC in the presence of a weak static defect
(Sec. ). In this context, we review
Bogoliubov's excitation spectrum and its associated Landau criterion
for superfluidity, followed by a detailed discussion on superfluidity
and related phenomena, such as quantized vortices, in
Sec. . We briefly touch on the subject of
synthetic gauge fields for neutral atoms, before investigating the
properties of BECs in periodic potentials created by optical lattices
(Sec. ). Finally, we combine the concepts of
synthetic gauge fields and optical lattices in
Section , where we show the main features of the
Harper-Hofstadter model, which was recently realized using
atomic gases.
Bose-Einstein condensation
BEC in noninteracting gas
In 1925, Albert Einstein (prompted by the earlier work of the indian
polyglot Satyendra Nath Bose) considered what would happen to a
non-interacting bosonic gas of non-relativistic particles in the
thermodynamic limit, as one lowers the temperature. He predicted the
phenomenon we now call Bose-Einstein condensation (BEC),
namely a phase transition to a new state of matter, in which a finite
fraction of all the particles would occupy the same single-particle
state. The transition occurs at fixed density below a critical
temperature or, alternatively, at fixed temperature, above a
critical density. In particular, if we take neutral particles in a
cubic box of volume , then they would predominantly occupy the
zero-wavevector state , and the critical temperature would
be
with the density and , being the particle
mass and Boltzmann's constant, respectively.
At its core, BEC is a paradigm of quantum statistical mechanics,
stemming from the indistinguishability of elementary particles and the
Bose-Einstein statistics that they obey. One can hand-wavingly deduce
the critical temperature (or critical density) where quantum
degeneracy would start playing a role in a many-body system, by
arguing that the thermal de Broglie wavelength should be comparable to
or greater than the inter-particle distance (which in our case is
on average) . Apart from the
numerical prefactor, we get the same answer as
Eq. eq:Tc3D. While one may argue that elementary massive
bosons do not exist, it is worth emphasizing that indistinguishability
only plays a role when there is a finite probability for exchange
processes to occur between the particles. In that sense, all
odd-isotope alkali atoms under relevant experimental conditions (see
below) effectively behave as bosons: their many-body wavefunction is
symmetric under the exchange of any two such atoms.
BEC in interacting system
Interestingly enough, BEC was considered by many at the time to be a
pathological behaviour of the non-interacting gas, which would resolve
once interactions were properly accounted for. In fact, it is well
known that the ideal Bose gas has infinite compresibility. This
pathology is cured by introducing a weak repulsive interaction between
bosons, a regime where BEC survives, as we will see next.
One-body density matrix
Following Leggett, we characterise each of the particles (assumed
spinless, for simplicity) by a position vector , with the label
running from 1 to . Any pure state of the (now interacting)
system - which can be also subjected to an external potential - can
be described at time by the many-body wavefunction
. Therefore, the most general state
of the system (also called mixed state) can be written as a
superposition of pure orthonormal states with different weights
. The single-particle density matrix
represents the probability
amplitude, at time , of finding a specific particle at position
, multiplied by the amplitude of finding it at ,
and averaged over the positions of all the other particles:
_1(,^,t) & N _s p_s d_2d_3d_N
_s^(,_2,,_N,t)_s(^,_2,,_N,t)
& = _i n_i(t) _i^(,t) _i(^,t)
where in the second line we have re-written the density matrix in
diagonal form, introducing its eigenvalues and eigenvectors
, which form a complete orthonormal set at any time (here
labels a good quantum number of the problem, i.e. momentum in a
translationally-symmetric situation).
Penrose-Onsager criterion for BEC
We are now ready to state the Penrose-Onsager criterion for
condensation, first formulated in 1956: if at any given time it is
possible to find a complete orthonormal basis of states of
such that one and only one of these states has an
eigenvalue of order (the rest being of order 1), then we say the
system exhibits BEC. One should note that this definition only applies
to "simple" BEC, as opposed to the "fragmented" case (of no
concern to us here), where two or more of the eigenvalues of the
one-body density matrix are of order .
Condensate wavefunction and the order parameter
We denote the single macroscopic eigenvalue of the density matrix by
, and its corresponding eigenfunction by
. is called the condensate
wavefunction and the particles occupying it the
condensate, while the ratio is the condensate
fraction. It is not necessarily true that , even at zero
temperature. Also note that, while behaves as a
single-particle Schrodinger wavefunction, it is generally not an
eigenfunction of the single-particle part of the Hamiltonian, or of
any other simple operator for that matter, other than
. Another useful quantity frequently found in the
literature is the so-called order parameter,
. We see that, while
is normalized to 1, will be normalized to .
No-go theorem for lower dimensionality
It is worth mentioning the existence of a theorem due to
Hohenberg , stating that, in the thermodynamic
limit, BEC cannot occur at a finite temperature in any system moving
freely in space in less than three dimensions, irespective of the
existence and/or sign of the interparticle interactions, as thermal
fluctuations would destroy the condensate. Note that this theorem,
however, only applies under equilibrium conditions, the nonequilibrium
case still being an open question. Furthermore, there is no general
proof that a realistic system of interacting bosonic particles must
show BEC, even at zero temperature - the solid phase of He
constitutes an obvious counter-example.
Experimental proof
Most gases, with the notable exception of He, are solids at the
densities and temperatures predicted by Eq. eq:Tc3D. That is
why it took no less than 70 years between Einstein's original paper
and the first experimental observation of BEC in an atomic gas. In
1995, the group of Eric Cornell and Carl Wieman succesfully condensed
a cloud of Rb atoms (closely followed by
the group of Wolfgang Ketterle at MIT with Na
atoms ), by first bringing the system to a
very low density, and then cooling it fast enough to prevent any
recombination processes that would have lead to the formation of the
solid phase. While other odd-isotope alkali elements, especially
Na or Li, are also routinely used in experiments, the
first non-alkali atom to be cooled into the BEC phase was hydrogen
H. Due to the extreme diluteness of these systems
( atoms/cm), the typical range of is from
nK to a few K. Achieving such ultra-low temperatures
stimulated the development of novel experimental techniques, such as
magnetic/laser trapping and evaporative cooling of atoms.
Diagnostic techniques
The very low densities of alkali gases also limit the range of
available diagnostic techniques. The most commonly employed method in
BEC experiments is optical absorption imaging, where one shines a
laser on the gas and detects the percentage of transmitted power. The
image is usually taken after removing the trap and allowing the gas to
expand. This gives information about the gas density as a function of
coordinates and time, with a spatial resolution of a few m. In
stark contrast to liquid He, density-related information seems
to be sufficient for most practical purposes.
Diluteness/weak interaction condition
Gross-Pitaevskii equation
Short history and utility of GP equation
As it turns out, many of the experimental results in ultracold gases
can be interpreted on the basis of a single equation for the
condensate wavefunction . This equation, first derived
in 1961 independently by Eugene Gross and Lev Pitaevskii, was
originally intended as a phenomenological description of quantum
vortices in the superfluid phase of liquid He, below the lambda
point. Since liquid helium is a strongly interacting system however,
the GP equation turned out to be much better suited to alkali gases.
Before giving the concrete formulation of the GP equation, we must
first explore the nature of the inter-atomic interactions.
s-wave scattering length
In dilute systems, the inter-atomic distance is on
the order of 1000 , while the range of the inter-atomic
potential, namely the extent of the last bound state of the van der
Waals interaction, is about 50-100 . As , the
probability of three-atom colissions is substantially diminished. This
justifies limiting ourselves to a binary (instead of three-body or
more) scattering problem: consider two atoms, separated by a relative
distance and interacting in three dimensions through a potential
. We can therefore decouple their center-of-mass motion from
their relative one and write a Schrodinger equation for the
scattering states .
For temperatures below , the thermal de Broglie wavelength
(as mentioned in Sec. ), meaning all
significantly occupied states will have small wavevectors,
. This directly translates to a low relative kinetic
energy, and hence small relative wavevectors, for the scattering
problem outlined above. However, we know from scattering theory that
the probability for two atoms, with relative angular momentum ,
of being separated by a distance is proportional to
, therefore essentially negligible in the limit
. That is of course, unless , meaning their
relative state is -wave, which is what we will assume from now
on. Since , one can use the asymptotic expression for
, which only depends on the scattering amplitude. At small
wavevectors, this amplitude can be safely replaced by the
-wave scattering length , which will encapsulate all
the interaction effects on macroscopic properties of the atomic gas.
One can now replace the two-body potential with an effective
interaction, , provided it gives the same
scattering length. The limit of small wavevectors prompts us to only
consider the lowest Fourier component of , equivalent
in real space to a contact interaction(Technically, one
should also include a regularizing part in order to remove any
divergencies of the wavefunction.)
, where we have introduced the
interaction coupling constant , whose value can be calculated using
the first-order Born approximation
The scattering length therefore becomes the small parameter of
the theory of weakly-interacting ultracold gases, and the validity of
the Born approximation rests on the following two conditions
k a_s & 1
a_s & ^-1/3
Eq. eq:diluteness-condition is called the "diluteness
condition", and it paves the way to various mean-field approaches,
such as the GP equation.
Derivation of Gross Pitaevskii equation
Formally, the GP equation corresponds to the lowest-order expansion in
of the more exact Bogoliubov theory. However, we will try to
give a hand-waving justification of it for the zero-temperature
case. At , all particles are in the condensate, therefore one
could neglect all inter-particle correlations and introduce the
simplest (Hartree-Fock) ansatz, expressing the ground state many-body
wavefunction in the symmetrized form
As mentioned in Sec. , the single-particle state
(now occupied by all the bosons) obeys a Schrodinger-like
equation, to which we must add the energy of the effective binary
interactions. In mean-field, these interactions contribute the
equivalent of a one-particle potential term proportional to
. Together with the kinetic part,
this results in the nonlinear equation(We have tacitly
assumed that is large enough, such that .)
where we have also included an external potential , normally
used to model harmonic trapping of the gas. Note that
Eq. eq:TDGP is valid for physics occuring over distances much
larger than the scattering length , which in turn must be smaller
than the typical range of the potential .
Caveats of TDGP
Eq. eq:TDGP is the time-dependent Gross-Pitaevskii (TDGP)
equation, a mean-field result where the condensate wavefunction
must be calculated self-consistently. It is important to
emphasize that the TDGP equation is also valid at nonzero temperatures
, provided that the density of non-condensed particles is
much smaller than the condensate density. In that case, the condensate
number is smaller (but still on the order of) the total particle
number . Finally, one must note that the nonlinearity of
Eq. eq:TDGP builds a bridge connecting BEC to nonlinear
optics, where a similar relation is used, under the name of
nonlinear Schrodinger equation.
Time-independent GP
In case the external potential does not depend explicitly on time,
the stationary solutions of Eq. eq:TDGP evolve with a trivial
phase factor . This yields a time-independent GP
equation for (we set from here on)
where is chemical potential of the gas, the energy required to
add one more particle to the system.(Technically, it is the
Lagrange multiplier associated to the conservation of particle
number , and can be shown to be very close to the actual
chemical potential in the thermodynamic limit. )
Linear response theory
Following loosely the formalism presented in Ref. , we now let be the steady state solution to the GP equation in
the time-independent trapping potential
with the GP Hamiltonian defined as
This Hamiltonian describes a bosonic condensate of particles
with contact interactions quantified by , and chemical potential
. Now consider adding a small time-dependent perturbation on top
of the trap, giving . We are
interested in the response of the condensate to this perturbation.
For weak perturbations, we can perform a linearization of the GP
equation Eq. eq:GP-atoms around the stationary solution
- an approach known in the literature as the "linear
response" formalism. The condensate wavefunction
evolves according to
We assume a small deviation of the wavefunction from its initial
steady state
such that we can expand Eq. eq:GP-atoms-evolution and keep only linear terms in and . We get
Note that Eq. eq:GP-atoms-lin is not strictly linear due to
the coupling of to . To restore
linearity, we consider the functions and
as being independent and write the linear system
where we have introduced the linear operator
and the source term . Note that
is a non-Hermitian operator!
We now consider the eigenvalue equation for the operator
with being the right eigenvector and its
corresponding eigenvalue
Similarly, we also introduce the left eigenvector, obeying
,
and the orthonormality condition
.
Notice that and are
connected by the unitary transformation(Note that this holds
as long as the Hamiltonian only contains real
terms.)
where is the third Pauli
matrix. We say that is -Hermitian, meaning
that one can define a new scalar product
with a
different signature, such that
. The operator is
usually called the metric operator, and, not suprisingly in our case,
it is the same as the one of the scalar Klein-Gordon equation. A
pseudo-Hermitian operator usually also posesses antilinear symmetries,
and as we will see below this is also the case for
. Interestingly, for operators with a real spectrum,
it can be shown that one can define another metric , which
guarantees a positive-definite inner product, or, in other words,
(provided of
course). This can be used to formulate a probabilistic quantum theory
for the new wave-functions and . For the general
theory and properties of pseudo-Hermitian operators, we point the
interested reader to Ref. .
Using Eq. eq:symmetry-1, we get the general form of the left
eigen-vectors as
with a normalization factor. We can chose
and group the eigenvalues of
into 3 families, according to the quantity
We therefore have: the "" family, corresponding to , the
"" family, such that and the "" family, with
.
We are now ready to write the completeness relation
Using Eq. eq:completeness, we can decompose any column vector
as(The modes in the "" family do not appear in this
expansion as their components live in the space orthogonal to the one
of our solution.)
l_1l_2 = _k "+" family [u_kl_1 - v_kl_2]u_kv_k
+ _k "-" family [v_kl_2 - u_kl_1]u_kv_k
There is now a further symmetry of that we can
exploit in our problem, a sort of time-reversal "spin"-flip
symmetry, namely
where , with
the first Pauli matrix and
the complex conjugation antilinear operator. This results in a
duality between the "" family with eigenvectors and
energy and the "" family with eigenvectors
and energy
.
We can now finally project Eq. eq:GP-atoms-system onto the
eigenvectors of . Using the above-mentioned duality and
Eq. eq:decomposition, we get
with the complex amplitudes satisfying
where we introduced
Cherenkov emission of Bogoliubov excitations
We now turn to applying the formalism developed in
Sec. to a concrete physical example, namely
a flowing condensate scattering against a static
defect . The BEC(We integrate all
density profiles along the direction, resulting in an effective
2-dimensional description.) is therefore in a state with
well-defined momentum, described by the plane wave
and a chemical potential . Since we have
no trap, , and Eq. eq:GP-atoms produces the
equation of state
where we have introduced the condensate density
.
We now introduce a weak perturbation in the form of a static localized
defect potential , which can represent
for instance a laser spot depleting a small area of the condensate, as
shown in Fig. .
Density profiles of an expanding BEC hitting a stationary defect
created by the repulsive potential of a blue-detuned laser beam. The
condensate has different speeds in the two panels, moving roughly
twice as fast in the right-panel. Notice the Mach cone formed behind
the defect, which gets narrower as the condesate moves faster. From
Ref. .
Top panels: Bogoliubov dispersion
Eq. eq:bogoliubov. The dotted lines indicate the
plane.
Middle panels: Locus of intersection of the 2D
dispersion with the plane. Green arrows are normal
to , while the dashed lines indicate the Cherenkov
cone.
Bottom panels: Real-space density modulation, with a
-defect at . Dashed lines show the Mach cone. Left
column panels are for , and right column for
. From Ref. .
Using Eq. eq:atom-MF, the GP Hamiltonian becomes
and the
source term
. We now
get the linear operator for our problem in the form
Notice that, due to the presence of the off-diagonal exponential
terms, does not commute with the momentum operator, which is
the generator of the spatial translation group. Luckily, however, we
can restore translational invariance by a simple unitary
transformation, as shown below.
Using the standard commutation relations, one can show that, for a
constant wavevector , the unitary operator(The hat
symbol denotes operators in the relevant Hilbert space.)
performs a translation in momentum space,
, with the
ket representing a single particle state with
wavevector such that
. Using the
definitions above, one can easily obtain the commutator
This allows us to rewrite the following expressions
T^(k_0)kT(k_0)& = k - k_0I
T(k_0)kT^(k_0)& = k + k_0I
We now recognize the two exponentials in Eq. eq:ourL as being
the real-space representation of and its
hermitian conjugate. This motivates us to define the following unitary
operator
such that a unitary transformation of our operator now restores
translational symmetry. Indeed, one can see that
where we have made use of Eqs. eq:products and we have written
in a base-independent representation. In the subspace of
momentum eigenstates , we can write the (right-)eigenvalue
equation corresponding to Eq. eq:translated-L as
where we have recovered the matrix representation of
Eq. eq:LGP, and introduced the notation
Here labels the 2 different eigenmodes, and we defined
the condensate speed .
Notice that the mode has only one eigenvector. However, one can
safely exclude it as this mode does not imply energy or momentum
transport. Excluding the point, one can then solve
Eq. eq:right-eigen, obtaining the celebrated Bogoliubov
excitation spectrum