ArticlePDF Available

Abstract

An expanding polariton condensate is investigated under pulsed nonresonant excitation with a small laser pump spot. Far above the condensation threshold we observe a pronounced increase in the dispersion curvature, with a subsequent linearization of the spectrum and strong luminescence from a ghost branch orthogonally polarized with respect to the linearly polarized condensate emission. Polarization of both branches is understood in terms of spin-dependent polariton-polariton scattering. The presence of the ghost branch has been confirmed in time-resolved measurements. The effects of disorder and dissipation in the photoluminescence of polariton condensates and their excitations are discussed.
1
Ghost Branch Photoluminescence From a Polariton Fluid Under Nonresonant Excitation
Maciej Pieczarka 1,*, Marcin Syperek 1, Łukasz Dusanowski 1, Jan Misiewicz 1, Fabian Langer 2,
Alfred Forchel 2, Martin Kamp 2,Christian Schneider 2, Sven Höfling 2,3, Alexey Kavokin 4,5,
Grzegorz Sęk 1
1 Laboratory for Optical Spectroscopy of Nanostructures, Department of Experimental Physics,
Wrocław University of Technology, Wybrzeże Wyspiańskiego 27, 50-370 Wrocław, Poland
2 Technische Physik, University of Würzburg and Wilhelm-Conrad-Röntgen-Research Center for
Complex Material Systems (RCCM), Am Hubland, D-97074 Würzburg, Germany
3 SUPA, School of Physics and Astronomy, University of St. Andrews, St. Andrews KY16 9SS,
United Kingdom
4 Spin Optics Laboratory, Saint Petersburg State University, 1, Ulianovskaya, 198504, St
Petersburg, Russia
5 Physics and Astronomy School, University of Southampton, Highfield, Southampton SO17 1BJ,
UK.
*corresponding author: maciej.pieczarka@pwr.edu.pl
PACS numbers: 71.36.+c, 67.10.-j, 71.35.Lk, 78.67.De
Abstract
An expanding polariton condensate is investigated under pulsed nonresonant excitation with a
small laser pump spot. Far above the condensation threshold we observe a pronounced increase in
the dispersion curvature with a subsequent linearization of the spectrum and strong luminescence
from a ghost branch orthogonally polarized with respect to the linearly polarized condensate
emission. The presence of the ghost branch has been confirmed in time-resolved measurements.
The dissipative and nonequilibrium effects in the photoluminescence of polariton condensates and
their excitations are discussed.
Exciton polaritons are composite bosons consisting of strongly coupled microcavity photons and
quantum well excitons. They are able to form a novel class of condensates (for recent reviews see,
e.g. [1], [2]). Despite their nonequilibrium and dissipative nature, where the polariton condensation
is governed by their amplified scattering and by the balance between the gain from the polariton
reservoir and the loss due to the escape of photons from the microcavity, they behave as
condensates of weakly interacting bosons. Collective phenomena like condensation [3], off
diagonal long range order [4], [5], topological excitations [6] or superfluid features in propagation
of polariton flows [7] have been demonstrated in such systems. A spectacular consequence of the
parametric scattering in polariton gases is appearance of non-parabolic scattering bands, including
normal branches (NBs) and ghost branches (GBs), the latter ones populated by the virtual off-
branch exciton-polaritons [7,8]. Excitations of polariton condensates are expected to be
2
characterized by linearized dispersions of the Bogoliubov-like spectra [9]. The fluid excitations at
low momenta are expected to behave as collective sound waves rather than single particles.
Recently, the linear dispersion NB and the GB of a resonantly pumped polariton condensate [10]
have been observed in a four-wave mixing experiment.
The experiment providing a direct access to the dispersion of excitations of spontaneously formed
condensates in polariton systems is photoluminescence (PL) under nonresonant excitation. The
nonresonant pumping can be realized either with a detuned laser or with the current
injection [11,12]. In the nonresonant pumping scheme, incoherent excitons are generated and relax,
subsequently feeding the reservoir and governing the dynamics of the system [13]. The relaxation
of created hot excitons involves multiple scattering processes, which destroy the coherence and
phase of the excitation, ensuring that these properties are not inherited by the condensate, in
contrast to the resonant excitation case. At the same time, the excessive noncondensed quasi-
particles build up an incoherent reservoir, causing additional decoherence [14] and forming a
repulsive potential [15], which shapes the condensate spatially and spectrally. This results in a high
sensitivity of polariton fluids to the size and profile of the optical pump beam. Moreover, it affects
significantly the excitation spectrum of the condensate [16]. Bogoliubov dispersions have been
reported in this excitation scheme, however, no fingerprints of the GB have been observed
yet [17], [18]. This fueled the debate on the possibility of observing the GB in nonresonant PL [19]
where the strong emission of the condensate occurs and that one being related to the NB can easily
mask the GB signal. The resonant experiment done by V. Kohnle et. al. in [10] is in stark contrast
to nonresonant pumping scheme, since the polariton condensate formation occurred in absence of
incoherent reservoir and the collective excitations were created by a resonant trigger pulse.
In this Letter, we report on the direct observation of a PL signal of the GB together with a
pronounced linearization of the dispersion of an expanding polariton condensate under nonresonant
excitation. We verify the observation via polarization- and time-resolved experiments, which
elucidate the origin of the observed features.
The investigated sample was a GaAs-based microcavity with embedded high-indium content
In0.3Ga0.7As quantum wells. The cavity consist of two Bragg mirrors (top and bottom) of alternating
GaAs and AlGaAs pairs of layers. The length of the cavity is equal to λ/2. The quality factor of the
cavity estimated from the reflectivity spectra taken far from the polariton anticrossing resonance
was around 1000. The system was in a strong coupling regime characterized by the normal mode
splitting of 7 meV. Experiments described in this letter were performed at negative photon-exciton
detunings of around -0.5 meV and -2 meV. Details on the experimental setup are given in the
supplementary material [20].
The non-linear properties of polariton condensates have been investigated using excitation power
dependent PL measurements both in real and momentum space. In our experiment the pumping
laser was focused to a diffraction limited Gaussian shape of about 2 µm in diameter. The PL
intensity as a function of the excitation power (average power measured outside of the cryostat) for
two linear and perpendicular polarization detections, co-linear and cross-linear to the polarization
axis of the polariton condensate, is shown in Fig. 1(a). The data is extracted from time-integrated
data, hence the values are averaged over many excitation pulses. At low excitation densities, the
3
lower polariton branch (LP) dispersion is formed as can be seen in Fig 1(b). With a further increase
of the excitation power, we observe a distinct nonlinear threshold in the intensity dependence on
pumping shown in Fig. 1(a), consistent with the formation of a flat dispersion characteristic of a
propagating polariton condensate [Fig. 1(c)] at the threshold. This kind of dispersion is predicted
[16] when exciting a polariton condensate by a small excitation spot. The condensation threshold
is also manifested by a decrease of the linewidth of the measured emission which drops down by a
factor of 5 at threshold, and by a blueshift of the emission peak of about 3 meV (not shown). The
energy shift of the propagating condensate created nonresonantly is composed of the kinetic energy
of polaritons, repulsive interactions within the condensate and a term coming from the interaction
of polaritons with pump-spot-induced reservoir of noncondensed quasi-particles [15,21]. Hence,
the condensate energy is shifted above the minimum of the bare cavity mode, while the strong
coupling is preserved.
The microcavity polaritons subject to a natural disorder potential which originates from the width
and composition fluctuations of the quantum wells and Bragg mirrors [22]. Much effort has been
directed to reduce the impact of disorder [21], by growing nearly perfect, defect-free cavities.
Nevertheless, even in the best quality samples the localization and scattering of polaritons by a
stationary disorder potential is inevitably present. We observe polariton trapping by the disorder
potential as one can see in Figure 2. Localization of polaritons in real space has been observed at
moderate pump densities in our experiment. Bright spots of the luminescence located outside the
pump area are visible in Fig. 2(a), in agreement with the flat parts of the polariton dispersion
observed in the reciprocal space [Fig. 2(c)] [24]. The most striking phenomena have occurred at
higher polariton densities. First, we observed an evident spreading of the polariton cloud to
distances much larger than the pump spot Fig. 2(b). This can be interpreted in terms of ballistic
flow of the polariton condensate pumped above the percolation threshold in a 2D disorder potential.
Furthermore, well above the threshold a pronounced linearization of the polariton emission and a
clear evidence of the GB emission at the negative energies (below the condensate emission energy)
have been observed, as presented in Fig. 2(d). The GB becomes visible at pump densities around
20Pth at different detunings and at different spot positions at the sample surface. The buildup of the
GB luminescence could be attributed to the off-branch multiple-scattering processes in the
polariton gas [7]. However, the intensity of GB is strong and comparable to the NB and we detected
a change in the dispersion curvature, indicating a collective phenomenon of mixing the NB and
GB [10]. The linear spectrum and the appearance of the GB are fingerprints of collective
Bogoliubov-like excitations of a propagating polariton fluid. Similar features can be induced
resonantly in a propagating condensate in the supersonic regime in the presence of obstacles which
disturb the flow [23], [25]. Here, one can observe interference ripples in the real space image [Fig.
2(d)], which is a signature of nonradial propagation of scattered polariton waves [26]. These
patterns are stable for many minutes as they are observed in a time-integrated picture, thus the
interference phenomena have a major contribution to the final integrated signal, as confirmed by
the reproducibility of the features after many pulses. The disorder in our sample is believed to be
the source of excitations in the expanding polariton fluid, behaving as a static scattering landscape.
Additionally, one has to note a slight anisotropy of the emission intensity distribution over the
dispersion branches, which reflects a non-uniform disorder potential in the real space. The latter
4
shapes both the coherent flow of polaritons and the scattered waves. The specific spectra and
interference patterns are determined by the position of the excitation spot on the sample surface.
Let us now discuss the polarization characteristics of emitted photons. One important property of
polaritons is their spin fine structure inherited from excitons and photons. The optically active LPs
have the spin projection
1S
on the growth axis direction. Due to the anisotropic spin
interactions the minimization of the system energy favors the superposition of spin up and spin
down polaritons at the condensation threshold, resulting in a linear polarization buildup [27].
Moreover, in an isotropic medium this is a signature of a spontaneous symmetry breaking induced
by the condensation, and the vector polarization can be considered as an order parameter [28], [27].
In our experiment the nonlinear threshold is accompanied by a rapid increase of the degree of linear
polarization (DOLP) around
0k
, presented in Fig. 3(a) (where
/DOLP I I I I

 
;
,
I
are
the intensities detected in two orthogonal linear polarizations). The condensate forms at each point
in the real space with the same direction of linear polarization vector. We can observe a small
polarization asymmetry of the polariton ground state which is why the polarization vector of the
condensate is always pinned to one of the crystal axes [29,30]. For higher pumping levels we
observe depinning of the polarization vector, manifested in a decrease of the total DOLP with
respect to the pump power [31]. This data is extracted from the time-integrated measurements, so
the measured DOLP values are averaged over hundreds of time evolutions, resulting in the lower
value of the DOLP compared to each individual realization [31]. However, it allows the effect of
the abovementioned depinning in the pump power dependent measurements to be visible. Now,
considering the polarization properties of the NB and GB, one should expect the NB to be polarized
as the condensate and the GB signal to be orthogonal to it. This is due to the physical origin of the
two coupled branches of excitations. While the NB is populated due to the non-zero effective
temperature of the polariton gas, the GB is populated due to the depletion of the condensate by
polariton-polariton scattering similarly to the off-branch scattered states observed at resonant
pumping in [7]. According to the polarization selection rules [32], scattering of linearly polarized
polaritons produces polaritons having an orthogonal linear polarization, preferentially. Indeed, our
polarization-resolved measurements have shown the GB to be orthogonally polarized with respect
to the condensate polarization axis. The results are shown in Figures 3(b), where signal from
separate GB and NB as a function of polarization angle is shown, and 3(e), where we plot a DOLP
map created from a direct overlap of the two dispersions from Fig. 3(c) and (d) (after correction to
the small polarization splitting of the order of 0.1 meV) and calculating the polarization degree for
each pixel of the recorded data. A large area of a positive DOLP is observed for the NB, which
comes from the background created by the full time evolution of the signal after the pump pulse.
More importantly, very distinct inversed polarization of the GB is clearly visible in the DOLP map.
It is worth noting that while the optical nonlinearities of the cavity itself can generate an analogue
of complex quantum fluid phenomena in the paraxial geometry [33], however the large value of
DOLP and linear polarization inversion for GB is a direct evidence of the polariton-polariton
scattering responsible for the population of GB. Clearly, this cannot be an effect of the linear
polarization splitting in the cavity, which could rather give a rise to the buildup of the circular
polarization of emission [34].
5
All the data discussed above has been obtained in the time-integrated measurements. The main
drawback of this kind of approach in experiments with pulsed excitation is a loss of the information
about the complex dynamics of the system after each excitation pulse, in particular, it leads to the
time-averaging of the blueshift of the polariton dispersion branch. This has been addressed in
Ref. [10], where the authors have observed an overestimated speed of sound in the time-integrated
picture as well as an asymmetry between slopes of the NB and GB. One can observe a similar
asymmetry in our experiment, where the NB has a greater slope than the GB - see Figs. 3(c-d). In
order to rule out the artifacts in our dispersions which might arise due to the time integration, we
have performed the time-resolved scanning of the measured dispersion of a propagating polariton
condensate at the detuning of 2 meV, and we studied in detail the dynamics of the polariton
luminescence after the pump pulse.
Approximately 50 ps after the pump pulse arrival we have observed a simultaneous and distinct
appearance of the positive and negative branches in the photoluminescence. A snapshot at 66 ps
after the pulse arrival is shown in Fig. 4(a). Clearly, they are not artifacts of time integration. One
can notice that the NB and GB have now comparable slopes in this time-resolved picture. However,
the extracted propagation velocity is still somewhat larger than the speed of sound calculated from
the condensate blueshift from Fig. 4(a): the velocity taken from the slope of the time-resolved
experiment is
1.95 /
slope
v m ps
and is greater than the velocity of
1.45 /
blue
v m ps
calculated
from the blueshift at
0k
, according to definition
, where U is the blueshift
(corresponding to mean field energy) and mpol is the LP effective mass. Confronting these velocities
to the values obtained from time-integrated series of a less photonic condensate corresponding to
the detuning of -0.5 meV, presented in Fig. 4(b), one can see even larger difference in both values.
Here, however, one has to take into account the overestimation of the slopes in the time-integrated
picture. The 𝑣𝑠𝑙𝑜𝑝𝑒 corresponds to the polariton mean field energy of
U
3.8 meV , being four
times larger than the observed temporal blueshift. One of the reasons might be technical, namely
the response resolution of the streak camera equipped with a monochromator is 12 ps. However,
the GB signal lasts for much longer, at least for 50 ps which excludes the apparatus effect. It has
to be noted, that the used excitation scheme creates a condensate with a nontrivial spatial
distribution of wave vectors [15]. Our case is far from the steady state of a static condensate and in
this experimental configuration the pump pulse creates moving condensate, characterized by the
outward coherent flow and disorder scattering of the polaritonic waves. This is why the condensate
at zero wave vector is more likely to be created via scattering on multiple defect centers [35]
resulting in lower blueshift than for a static condensate. We can also speculate on the spatially
varying local Doppler shift of created excitations which might be responsible for the observed
distinct slopes, adding additionally the local condensate velocity to the slope of the dispersion.
This, however, recalls for a further theoretical analysis.
The observed emission of the condensate and the GB are present at the energy very close or even
above the energy of the cavity mode in our sample, giving rise to the question if these phenomena
occur still in the strong coupling regime. In the time-resolved spectrum right after the pump pulse
arrival, we observe a strong emission corresponding to the bare cavity mode at the time scale below
6
the temporal resolution of our setup as shown in Fig. 4(c). All the recorded dispersion features
including the linearized NB and the orthogonally polarized GB signal occur later in time, where
the weak coupling lasing has completely vanished. We conclude that these spectral features are
characteristics of a polariton condensate formed in the strong coupling regime. In fact, in the
time-resolved measurements we observe a transition from weak to strong coupling similar to the
one reported in Ref. [36]. We believe that we were able to observe very clearly the renormalization
of the NB and the buildup of the GB partly due to the relatively modest quality factor of our sample,
which was low enough to enable efficient extraction of photons outside the cavity and high enough
to preserve the conditions for the polariton condensate formation. Moreover, the intrinsic disorder
of the cavity enhanced the elastic scattering favorable for the condensate formation.
In conclusion, we report on the first observation of a polariton ghost branch in a photoluminescence
experiment under nonresonant excitation. Even though the system in our approach was far from
equilibrium and from the steady state regime, we have observed a distinct renormalization of the
polariton dispersion characterized by the Bogoliubov-like normal and ghost branches. The origin
of the observed dispersion branches has been identified based on the observed orthogonal linear
polarizations of the branches and on the time-resolved dispersions showing temporal symmetric
slopes for the normal and ghost branches of the polariton condensate excitations.
Acknowledgments
We would like to thank Iacopo Carusotto for critical reading of this manuscript and for valuable
discussions. The authors acknowledge the financial support from the bilateral project of Deutsche
Forschungsgemeinschaft (project named LIEPOLATE) and Polish Ministry of Science and Higher
Education (project No. DPN/N99/DFG/2010). The experiment is partially performed within the
NLTK infrastructure, Project No. POIG. 02.02.00-003/08-00. AK acknowledges the financial
support from the Russian Ministry of Education and Science (Contract No.11.G34.31.0067).
7
References
[1] T. Byrnes, N. Y. Kim, and Y. Yamamoto, Nat. Phys. 10, 803 (2014).
[2] I. Carusotto and C. Ciuti, Rev. Mod. Phys. 85, 299 (2013).
[3] J. Kasprzak, M. Richard, S. Kundermann, a Baas, P. Jeambrun, J. M. J. Keeling, F. M.
Marchetti, M. H. Szymańska, R. André, J. L. Staehli, V. Savona, P. B. Littlewood, B.
Deveaud, and L. S. Dang, Nature 443, 409 (2006).
[4] J. Fischer, I. G. Savenko, M. D. Fraser, S. Holzinger, S. Brodbeck, M. Kamp, I. A.
Shelykh, C. Schneider, and S. Höfling, Phys. Rev. Lett. 113, 203902 (2014).
[5] F. Manni, K. G. Lagoudakis, R. André, M. Wouters, and B. Deveaud, Phys. Rev. Lett.
109, 150409 (2012).
[6] G. Nardin, G. Grosso, Y. Léger, B. Piȩtka, F. Morier-Genoud, and B. Deveaud-Plédran,
Nat. Phys. 7, 635 (2011).
[7] P. Savvidis, C. Ciuti, J. Baumberg, D. Whittaker, M. Skolnick, and J. Roberts, Phys. Rev.
B 64, 075311 (2001).
[8] J. Zajac and W. Langbein, arXiv1210.1455 (2012).
[9] L. Pitaevskii and Stringari S., Bose-Einstein Condensation (Oxford University Press, New
York, 2003).
[10] V. Kohnle, Y. Léger, M. Wouters, M. Richard, M. T. Portella-Oberli, and B. Deveaud-
Plédran, Phys. Rev. Lett. 106, 255302 (2011).
[11] C. Schneider, A. Rahimi-Iman, N. Y. Kim, J. Fischer, I. G. Savenko, M. Amthor, M.
Lermer, A. Wolf, L. Worschech, V. D. Kulakovskii, I. A. Shelykh, M. Kamp, S.
Reitzenstein, A. Forchel, Y. Yamamoto, and S. Höfling, Nature 497, 348 (2013).
[12] P. Bhattacharya, T. Frost, S. Deshpande, M. Z. Baten, A. Hazari, and A. Das, Phys. Rev.
Lett. 112, 236802 (2014).
[13] F. Tassone, C. Piermarocchi, and V. Savona, Phys. Rev. B 56, 7554 (1997).
[14] J. Schmutzler, T. Kazimierczuk, Ö. Bayraktar, M. Aßmann, M. Bayer, S. Brodbeck, M.
Kamp, C. Schneider, and S. Höfling, Phys. Rev. B 89, 115119 (2014).
[15] M. Wouters, I. Carusotto, and C. Ciuti, Phys. Rev. B 77, 115340 (2008).
[16] M. Wouters and I. Carusotto, Phys. Rev. Lett. 99, 140402 (2007).
8
[17] S. Utsunomiya, L. Tian, G. Roumpos, C. W. Lai, N. Kumada, T. Fujisawa, M. Kuwata-
Gonokami, A. Löffler, S. Höfling, A. Forchel, and Y. Yamamoto, Nat. Phys. 4, 700
(2008).
[18] M. Assmann, J.-S. Tempel, F. Veit, M. Bayer, A. Rahimi-Iman, A. Löffler, S. Höfling, S.
Reitzenstein, L. Worschech, and A. Forchel, Proc. Natl. Acad. Sci. U. S. A. 108, 1804
(2011).
[19] T. Byrnes, T. Horikiri, N. Ishida, M. Fraser, and Y. Yamamoto, Phys. Rev. B 85, 075130
(2012)
[20] See Supplementary Meterial.
[21] P. Cilibrizzi, A. Askitopoulos, M. Silva, F. Bastiman, E. Clarke, J. M. Zajac, W. Langbein,
and P. G. Lagoudakis, Appl. Phys. Lett. 105, 191118 (2014).
[22] V. Savona, J. Phys. Condens. Matter 19, 295208 (2007).
[23] I. Carusotto and C. Ciuti, Phys. Rev. Lett. 93, 166401 (2004).
[24] D. Sanvitto, A. Amo, L. Viña, R. André, D. Solnyshkov, and G. Malpuech, Phys. Rev. B
80, 045301 (2009).
[25] A. Amo, J. Lefrère, S. Pigeon, C. Adrados, C. Ciuti, I. Carusotto, R. Houdré, E. Giacobino,
and A. Bramati, Nat. Phys. 5, 805 (2009).
[26] G. Christmann, G. Tosi, N. G. Berloff, P. Tsotsis, P. S. Eldridge, Z. Hatzopoulos, P. G.
Savvidis, and J. J. Baumberg, Phys. Rev. B 85, 235303 (2012).
[27] F. Laussy, I. Shelykh, G. Malpuech, and A. Kavokin, Phys. Rev. B 73, 035315 (2006).
[28] H. Ohadi, E. Kammann, T. C. H. Liew, K. G. Lagoudakis, A. V. Kavokin, and P. G.
Lagoudakis, Phys. Rev. Lett. 109, 016404 (2012).
[29] J. Kasprzak, R. André, L. Dang, I. A. Shelykh, A. V. Kavokin, Y. Rubo, K. Kavokin, and
G. Malpuech, Phys. Rev. B 75, 045326 (2007).
[30] Ł. Kłopotowski, M. D. Martín, A. Amo, L. Viña, I. A. Shelykh, M. M. Glazov, G.
Malpuech, A. V. Kavokin, and R. André, Solid State Commun. 139, 511 (2006).
[31] J. Levrat, R. Butté, T. Christian, M. Glauser, E. Feltin, J.-F. Carlin, N. Grandjean, D. Read,
A. V. Kavokin, and Y. G. Rubo, Phys. Rev. Lett. 104, 166402 (2010).
[32] T. Ostatnický and A. V. Kavokin, Superlattices Microstruct. 47, 39 (2010).
[33] J. Scheuer and M. Orenstein, Science 285, 230 (1999).
9
[34] P. Cilibrizzi, H. Ohadi, T. Ostatnicky, A. Askitopolous, W. Langbein, and P. Lagoudakis,
Phys. Rev. Lett. 113, 103901 (2014).
[35] M. Wouters and I. Carusotto, Phys. Rev. Lett. 105, 020602 (2010).
[36] E. Kammann, H. Ohadi, M. Maragkou, A. V. Kavokin, and P. G. Lagoudakis, New J.
Phys. 14, 105003 (2012).
Figure captions
Fig. 1. a) Power dependent intensity curve of two linear polarizations, co-linear to the condensate
axis (blue circles) and cross-linear to the condensate (orange circles). Time integrated dispersions
below-(a), and slightly above the threshold-(b). LP branch (white solid line), blueshifted (blue
dashed) and bare cavity mode (white dashed) dispersions are shown. The color scale is linear.
Measurements for detuning -0.5 meV.
Fig. 2. Localization of polaritons at moderate densities in the momentum (a) and real space (c).
High pumping conditions with onset of linear dispersions (b) and the extended polariton cloud (d).
Line description is the same as in Fig. 1. The pump spot location is indicated as “x” in (b) and (d).
The color scale is linear.
Fig. 3. (a) DOLP as a function of the pumping power. (b) Polarization of separate NB and GB
signal extracted from the PL map measurements. The values are normalized. Co-linear (c) and
cross-linear (d) detection with respect to the condensate polarization and corresponding DOLP map
(e) of dispersion at the detuning of -0.5 meV. The DOLP scale is in the inset of (e) [-0.6;0.6]. The
intensity color scale is linear in (c) and (d).
Fig. 4. (a) Dispersion snapshot at 66 ps of the time-resolved evolution of the polariton dispersion
at the detuning of -2 meV . Cavity mode and LP are shown by white dashed and white solid lines,
respectively. The violet dotted lines are guides to the eye to highlight the similar slopes of the linear
dispersion part. The color map scales are logarithmic. (b) Velocities extracted from the slope of
time integrated data (orange circles) and sound speed calculated from dispersion blueshift (green
circles). (c) Snapshot at the ultrashort times after the pulse arrival, presenting photon lasing in the
weak coupling regime. The color scale is logarithmic.
10
Fig. 1
110 100
0,01
1
100
10000
Intensity (a.u.)
Power (mW)
a)
b) c)
11
Fig. 2
12
Fig. 3
13
Fig. 4
... However, these predictions have yet to be confirmed experimentally, and indeed it has recently been shown that any such modifications occur at momenta that are too low to be directly resolved in highquality microcavities [34]. Furthermore, while the observed shape of the Bogoliubov spectrum in experiments has so far appeared to be consistent with conventional Bogoliubov theory [34][35][36][37][38][39][40][41][42][43], there are notable unexplained observations such as the recently measured occupations of the excitation branches [41]. ...
... We can see that it is composed of two branches symmetric with respect to the condensate chemical potential (which is defined as the origin here). We denote the positive one as the normal branch (NB) and the negative one as the ghost branch (GB) [37]. Note that the ghost branch is not a real excitation branch of the system: As depicted in Fig. 3(b), it corresponds to the emitted photons associated with the creation of real Bogoliubov excitations in the polariton system, and as such it is related to the photoluminescence process itself (i.e., to the photons escaping the cavity). ...
... II D and previous theoretical works [29,31,32,60], the collective excitation spectrum of a polariton condensate is visible in the photoluminescence from the cavity. Indeed, signatures of collective excitations have been reported in experiments with nonresonant pumping [34,35,37,41], and in the different regime of coherent excitation [36,38,40]. In practice, however, the strong signal of the k = 0 state dominates the luminescence and saturates the detection system, making it challenging to detect the weak signal from the excitation branches. ...
Article
The classic Bogoliubov theory of weakly interacting Bose gases rests upon the assumption that nearly all the bosons condense into the lowest quantum state at sufficiently low temperatures. Here we develop a generalized version of Bogoliubov theory for the case of a driven-dissipative exciton-polariton condensate with a large incoherent uncondensed component, or excitonic reservoir. We argue that such a reservoir can consist of both excitonic high-momentum polaritons and optically dark superpositions of excitons across different optically active layers, such as multiple quantum wells in a microcavity. In particular, we predict interconversion between the dark and bright (light-coupled) excitonic states that can lead to a dynamical equilibrium between the condensate and reservoir populations. We show that the presence of the reservoir fundamentally modifies both the energy and the amplitudes of the Bogoliubov quasiparticle excitations due to the non-Galilean-invariant nature of polaritons. Our theoretical findings are supported by our experiment, where we directly detect the Bogoliubov excitation branches of an optically trapped polariton condensate in the high-density regime. By analyzing the measured occupations of the excitation branches, we extract the Bogoliubov amplitudes across a range of momenta and show that they agree with our generalized theory.
... GPE-based approaches describe reasonably well the dynamics of macroscopically-populated polariton condensates. At the same time, the Bogoliubov spectra derived for the excitations on top of such driven-dissipative GPEs reveal either gapped or diffusive real part of the dispersion [11-13], which does not correspond to experimental observations to date [24][25][26][27][28]. A combined GPE-Bolzmann equation for the condensate [14] recovers the gapless, equilibrium-like linearization of the excitation spectrum at small momenta, with the sound velocity defined by the condensate density n 0 : c s = gX 4 0 n 0 /m LP , where m LP is the lower polariton effective mass, g the exciton-exciton interaction strength and X p the exciton Hopfield coefficient at momentum p. ...
... GPE-based approaches describe reasonably well the dynamics of macroscopically-populated polariton condensates. At the same time, the Bogoliubov spectra derived for the excitations on top of such driven-dissipative GPEs reveal either gapped or diffusive real part of the dispersion [11][12][13], which does not correspond to experimental observations to date [24][25][26][27][28]. A combined GPE-Bolzmann equation for the condensate [14] recovers the gapless, equilibrium-like linearization of the excitation spectrum at small momenta, with the sound velocity defined by the condensate density n 0 : c s = gX 4 0 n 0 /m LP , where m LP is the lower polariton effective mass, g the exciton-exciton interaction strength and X p the exciton Hopfield coefficient at momentum p. ...
... [15][16][17] for resonant excitation. In nonresonant pumping scheme experiments [25][26][27][28], however, the slope of this linear part was shown to depend on the chemical potential (observed condensate blueshift) controversially for different temperatures, and even to deviate from the square-root scaling law [29]. This controversy indicates that the effects of finite temperature and the optically-dark reservoir on the elementary excitations of polariton condensates are poorly understood. ...
Preprint
Exciton-polaritons in an optical microcavity can form a macroscopically coherent state despite being an inherently driven-dissipative system. In comparison with equilibrium bosonic fluids, polaritonic condensates possess multiple peculiarities that make them behave differently from well-known textbook examples. One such peculiarity is the presence of dark excitons which are created by the pump together with optically-active particles. They can considerably affect the spectrum of elementary excitations of the condensate and hence change its superfluid properties. Here, we theoretically analyze the influence of the bright and dark "reservoir" populations on the sound velocity $c_s$ of incoherently-driven polaritons. Both pulsed and continuous-wave pumping schemes characterized by essentially different condensate-to-reservoir ratio are considered. We show that in the renormalized expression for the sound velocity the dark exciton contribution is dominant compared to that of thermal bright reservoir, and leads to considerable lowering of $c_s$ and to its deviation from the square-root-like behavior on the system's chemical potential (measurable condensate blueshift).
... However, these predictions have yet to be confirmed experimentally, and indeed it has recently been shown that any such modifications occur at momenta that are too low to be directly resolved in high-quality microcavities [34]. Furthermore, while the observed shape of the Bogoliubov spectrum in experiments has so far appeared to be consistent with conventional Bogoliubov theory [34][35][36][37][38][39][40][41][42][43], there are notable unexplained observations such as the recently measured occupations of the excitation branches [41]. ...
... We can see that it is composed of two branches symmetric with respect to the condensate chemical potential (which is defined as the origin here). We denote the positive one as the normal branch (NB) and the negative one as the ghost branch (GB) [37]. Note that the ghost branch is not a real excitation branch of the system: As depicted in Fig. 3(b), it corresponds to the emitted photons associated with the creation of real Bogoliubov excitations in the polariton system, and as such it is related to the photoluminescence process itself (i.e., to the photons escaping the cavity). ...
... II D and previous theoretical works [29,31,32,62], the collective excitation spectrum of a polariton condensate is visible in the photoluminescence from the cavity. Indeed, signatures of collective excitations have been reported in experiments with nonresonant pumping [34,35,37,41], and in the different regime of coherent excitation [36,38,40]. In practice, however, the strong signal of the k = 0 state dominates the luminescence and saturates the detection system, making it challenging to detect the weak signal from the excitation branches. ...
Preprint
Full-text available
The classic Bogoliubov theory of weakly interacting Bose gases rests upon the assumption that nearly all the bosons condense into the lowest quantum state at sufficiently low temperatures. Here we develop a generalized version of Bogoliubov theory for the case of a driven-dissipative exciton-polariton condensate with a large incoherent uncondensed component, or excitonic reservoir. We argue that such a reservoir can consist of both excitonic high-momentum polaritons and optically dark superpositions of excitons across different optically active layers, such as multiple quantum wells in a microcavity. In particular, we predict interconversion between the dark and bright (light-coupled) excitonic states that can lead to a dynamical equilibrium between the condensate and reservoir populations. We show that the presence of the reservoir fundamentally modifies both the energy and the amplitudes of the Bogoliubov quasiparticle excitations due to the non-Galilean-invariant nature of polaritons. Our theoretical findings are supported by our experiment, where we directly detect the Bogoliubov excitation branches of an optically trapped polariton condensate in the high-density regime. By analyzing the measured occupations of the excitation branches, we extract the Bogoliubov amplitudes across a range of momenta and show that they agree with our generalized theory.
... We achieve this by using an experimental technique where a strong resonant laser pump is employed to excite up-converted excitations, which are observed directly in the transmission spectrum, as schematically shown in Fig. 1(a). An analogous method was used to observe collective Bogoliubov excitations [28][29][30][31][32]. In our case, photon-lasing threshold is absent and the coherence is directly inherited from the pumping laser. ...
Article
Full-text available
In this work, we observe natural exceptional points in the excitation spectrum of an exciton–polariton system by optically tuning the light–matter interactions. The observed exceptional points do not require any spatial or polarization degrees of freedom and result solely from the transition from weak to strong light–matter coupling. It was demonstrated that they do not coincide with the threshold for photon lasing, confirming previous theoretical predictions [Phys. Rev. Lett. 122, 185301 (2019)PRLTAO0031-900710.1103/PhysRevLett.122.185301, Optica 7, 1015 (2020)OPTIC82334-253610.1364/OPTICA.397378]. Using a technique where a strong coherent laser pump induces up-converted excitations, we encircle the exceptional point in the parameter space of coupling strength and particle momentum. Our method of local optical control of light–matter coupling paves the way to the investigation of fundamental phenomena, including dissipative phase transitions and non-Hermitian topological states.
... At the same time, the momentum-energy dispersion on top of the condensate becomes linearized in accordance with the textbook Bogoliubov prediction 20 , therefore giving origin to superfluidity as follows from the Landau criterion. Multiple experiments with cavity exciton-polaritons have provided evidence of such linearization [21][22][23][24][25] despite the drivendissipative nature of the system suggesting the diffusive character of the Bogoliubov dispersion 26,27 with a zero real part in the region of small momenta. Precise observation of the shape of the excitations spectrum due to the thermal and quantum depletion of the macroscopically occupied ground state is also possible, but requires substantial momentum-space filtering covering a much brighter signal from the condensate 28,29 or using refined interferometric techniques 25 . ...
Article
Full-text available
Spectra of low-lying elementary excitations are critical to characterize properties of bosonic quantum fluids. Usually these spectra are difficult to observe, due to low occupation of non-condensate states compared to the ground state. Recently, low-threshold Bose-Einstein condensation was realised in a symmetry-protected bound state in the continuum, at a saddle point, thanks to coupling of this electromagnetic resonance to semiconductor excitons. While it has opened the door to long-living polariton condensates, their intrinsic collective properties are still unexplored. Here we unveil the peculiar features of the Bogoliubov spectrum of excitations in this system. Thanks to the dark nature of the bound-in-the-continuum state, collective excitations lying directly above the condensate become observable in enhanced detail. We reveal interesting aspects, such as energy-flat parts of the dispersion characterized by two parallel stripes in photoluminescence pattern, pronounced linearization at non-zero momenta in one of the directions, and a strongly anisotropic velocity of sound.
... We achieve this by using an experimental technique where a strong resonant laser pump is used to excite up-converted excitations, which are observed directly in the transmission spectrum, as schematically shown in Fig. 1(a). An analogous method was used to observe collective Bogoliubov excitations [28][29][30][31][32]. In our case, photon lasing threshold is absent and the coherence is directly inherited from the pumping laser. ...
Preprint
Full-text available
We observe natural exceptional points in the excitation spectrum of an exciton-polariton system by optically tuning the light-matter interactions. The observed exceptional points do not require any spatial or polarization degrees of freedom and result solely from the transition from weak to strong light-matter coupling. We demonstrate that they do not coincide with the threshold for photon lasing, confirming previous theoretical predictions [1, 2]. Using a technique where a strong coherent laser pump induces up-converted excitations, we encircle the exceptional point in the parameter space of coupling strength and particle momentum. Our method of local optical control of light-matter coupling paves the way to investigation of fundamental phenomena including dissipative phase transitions and non-Hermitian topological states.
... Firstly, because the polariton BEC is open-dissipative, the excitations of an homogeneous polariton condensate exhibit exotic properties. For instance, the linear excitations are provided by the diffusive Goldstone modes [22][23][24][25], with observable ghost branches of Bogoliubov excitations [26]. These have already triggered questions and studies on the definition of superfluidity and the characteristic observables in a nonequilibrium context, e.g., an extension of the standard Landau critical velocity has been proposed [11,12,[27][28][29][30][31]. ...
Article
Full-text available
Exploring the dynamics of a mobile impurity immersed in field excitations is challenging, as it requires to account for the entanglement between the impurity and the surrounding excitations. To this end, the impurity’s effective mass has to be considered as finite, rather than infinite. Here, we theoretically investigate the interaction between a finite-mass impurity and a dissipative soliton representing nonlinear excitations in the polariton Bose–Einstein condensate (BEC). Using the Lagrange variational method and the open-dissipative Gross–Pitaevskii equation, we analytically derive the interaction phase diagram between the impurity and a dissipative bright soliton in the polariton BEC. Depending on the impurity mass, we find the dissipative soliton colliding with the impurity can transmit through, get trapped, or be reflected. This work opens a new perspective in understanding the impurity dynamics when immersed in field excitations, as well as potential applications in information processing with polariton solitons.
... As such, the excitations of an homogeneous open-dissipative polariton condensate exhibit exotic properties. The linear excitations are provided by the diffusive Goldstone modes [20][21][22][23], with observable ghost branches of Bogoliubov excitations [24]. This have already prompted questions and studies on the definition of superfluidity and the characteristic observables in a nonequilibrium context, e.g., extension of the standard Landau critical velocity has been proposed [11,12,[25][26][27][28][29]. ...
Preprint
Full-text available
An important challenge in exploring the dynamics of a mobile impurity immersed in the field excitations comes from the necessity to include the entanglement between the impurity and the surrounding excitations. To address this challenge, the impurity's effective mass has to be assumed as finite, rather than infinite. Here we theoretically investigate how a finite-mass impurity interacts with a nonlinear excitation (i.e., soliton) in the polariton Bose-Einstein condensate (BEC), which is intrinsically non-equilibrium. Using the Lagrange variational method and the open-dissipative Gross-Pitaevskii equation, we analytically derive the interaction phase diagram between a bright soliton and an impurity in a polariton BEC. We show that when the soliton collide with the impurity, it can transmit through, be trapped, or reflected. Our work goes beyond prior researches in the context of equilibrium systems, and opens a new perspective toward understanding the nonequilibrium dynamics of a mobile impurity immersed in the nonliear field excitations.
Article
Harmonic oscillators belong to the most fundamental concepts in physics and are central to many current research fields such as circuit QED, cavity optomechanics, and photon pressure systems. Here, we engineer a microwave mode in a superconducting LC circuit that mimics the dynamics of a negative mass oscillator, and couple it via photon pressure to a second low-frequency circuit. We demonstrate that the effective negative mass dynamics lead to an inversion of dynamical backaction and to sideband cooling of the low-frequency circuit by a blue-detuned pump field, which can be intuitively understood by the inverted energy ladder of a negative mass oscillator.
Article
Exciton-polaritons in an optical microcavity can form a macroscopically coherent state despite being an inherently driven-dissipative system. In comparison with equilibrium bosonic fluids, polaritonic condensates possess multiple peculiarities that make them behave differently from well-known textbook examples. One such peculiarity is the presence of dark excitons which are created by the pump together with optically active particles. They can considerably affect the spectrum of elementary excitations of the condensate and hence change its superfluid properties. Here, we theoretically analyze the influence of the bright and dark “reservoir” populations on the sound velocity cs of incoherently driven polaritons. Both pulsed and continuous-wave pumping schemes characterized by essentially different condensate-to-reservoir ratios are considered. We show that the dark exciton contribution leads to considerable lowering of cs and to its deviation from the square-root-like behavior on the system's chemical potential (measurable condensate blueshift). Importantly, our model allows us to unambiguously define the density of dark excitons in the system by experimentally tracking cs against the condensate blueshift and fitting the dependence at a given temperature.
Article
Full-text available
Recently a new type of system exhibiting spontaneous coherence has emerged -- the exciton-polariton condensate. Exciton-polaritons (or polaritons for short) are bosonic quasiparticles that exist inside semiconductor microcavities, consisting of a superposition of an exciton and a cavity photon. Above a threshold density the polaritons macroscopically occupy the same quantum state, forming a condensate. The lifetime of the polaritons are typically comparable to or shorter than thermalization times, making them possess an inherently non-equilibrium nature. Nevertheless, they display many of the features that would be expected of equilibrium Bose-Einstein condensates (BECs). The non-equilibrium nature of the system raises fundamental questions of what it means for a system to be a BEC, and introduces new physics beyond that seen in other macroscopically coherent systems. In this review we focus upon several physical phenomena exhibited by exciton-polariton condensates. In particular we examine topics such as the difference between a polariton BEC, a polariton laser, and a photon laser, as well as physical phenomena such as superfluidity, vortex formation, BKT (Berezinskii-Kosterlitz-Thouless) and BCS (Bardeen-Cooper-Schrieffer) physics. We also discuss the physics and applications of engineered polariton structures.
Article
Full-text available
The investigation of intrinsic interactions in polariton condensates is currently limited by the photonic disorder of semiconductor microcavity structures. Here, we use a strain compensated planar GaAs/AlAs0.98P0.02 microcavity with embedded InGaAs quantum wells having a reduced cross-hatch disorder to overcome this issue. Using real and reciprocal space spectroscopic imaging under non-resonant optical excitation, we observe polariton condensation and a second threshold marking the onset of photon lasing, i.e., the transition from the strong to the weak-coupling regime. Condensation in a structure with suppressed photonic disorder is a necessary step towards the implementation of periodic lattices of interacting condensates, providing a platform for on chip quantum simulations.
Article
Full-text available
In this work, we combine a systematic experimental investigation of the power- and temperature-dependent evolution of the spatial coherence function, $g^{(1)}(r)$, in a one-dimensional exciton-polariton channel with a modern microscopic numerical theory based on a stochastic master equation approach. The spatial coherence function $g^{(1)}(r)$ is extracted via the high-precision Michelson interferometry, which allows us to demonstrate that in the regime of non-resonant excitation, the dependence $g^{(1)}(r)$ reaches a saturation value with a plateau, determined by the intensity of pump and effective temperature of the crystal lattice. The theory, which for the first time allows treating incoherent excitation in a stochastic frame, matches the experimental data with excellent qualitative and quantitative agreement. This allows us to verify the prediction that the decay of off diagonal long range order can be almost fully suppressed in one-dimensional condensate systems.
Article
Full-text available
One-dimensional polariton condensates (PoCos) in a photonic wire are generated through non-resonant laser excitation, by which also a reservoir of background carriers is created. Interaction with this reservoir may affect the coherence of the PoCo, which is studied here by injecting a condensate locally and monitoring the coherence along the wire. While the incoherent reservoir is mostly present within the excitation laser spot, the condensate can propagate ballistically through the wire. Photon correlation measurements show that far from the laser spot the second order correlation function approaches unity value, as expected for the coherent condensed state. When approaching the spot, however, the correlation function increases up to values of 1.2 showing the addition of noise to the emission due to interaction with the reservoir. This finding is substantiated by measuring the first order coherence by a double slit experiment, which shows a reduced visibility of interference at the excitation laser spot.
Article
Full-text available
We investigate the propagation and scattering of polaritons in a planar GaAs microcavity in the linear regime under resonant excitation. The propagation of the coherent polariton wave across an extended defect creates phase and intensity patterns with identical qualitative features previously attributed to dark and half-dark solitons of polaritons. We demonstrate that these features are observed for negligible nonlinearity (i.e. polariton-polariton interaction) and are therefore not sufficient to identify dark- and half-dark solitons. A linear model based on the Maxwell equations is shown to reproduce the experimental observations.
Article
Full-text available
We report on a real-time observation of the crossover between photon and exciton-polariton lasing in a semiconductor microcavity. Both lasing phases are observed at different times after a high-power excitation pulse. Energy-, time- and angle-resolved measurements allow for the transient characterization of carrier distribution and effective temperature. We find signatures of Bose–Einstein condensation, namely macroscoping occupation of the ground state and narrowing of the linewidth in both lasing regimes. The Bernard–Douraffourgh condition for inversion was tested and the polariton laser as well as the photon laser under continuous wave excitation were found to operate at estimated densities below the theoretically predicted inversion threshold.
Article
Full-text available
Zinc-blende semiconductor heterostructures grown in the [001] direction with a small lattice mismatch accommodate stress by developing a cross-hatch dislocation pattern. In GaAs based planar microcavities grown by molecular beam epitaxy, this pattern creates a potential landscape for exciton-polaritons, causing scattering and localization. We report here on suppressing the cross-hatch by introducing strain-compensating AlP layers into the center of the low index AlAs layers of the distributed Bragg reflectors. We observe a reduction of the cross-hatch dislocation density by at least one order of magnitude for 1.1 nm thick AlP layers, which correspond to an effective AlAs0.985P0.015 low index layer. These compensated structures show a remaining polariton disorder potential in the 10 μeV range.
Article
Full-text available
Conventional semiconductor laser emission relies on stimulated emission of photons, which sets stringent requirements on the minimum amount of energy necessary for its operation. In comparison, exciton-polaritons in strongly coupled quantum well microcavities can undergo stimulated scattering that promises more energy-efficient generation of coherent light by 'polariton lasers'. Polariton laser operation has been demonstrated in optically pumped semiconductor microcavities at temperatures up to room temperature, and such lasers can outperform their weak-coupling counterparts in that they have a lower threshold density. Even though polariton diodes have been realized, electrically pumped polariton laser operation, which is essential for practical applications, has not been achieved until now. Here we present an electrically pumped polariton laser based on a microcavity containing multiple quantum wells. To prove polariton laser emission unambiguously, we apply a magnetic field and probe the hybrid light-matter nature of the polaritons. Our results represent an important step towards the practical implementation of polaritonic light sources and electrically injected condensates, and can be extended to room-temperature operation using wide-bandgap materials.
Article
Room temperature electrically pumped inversionless polariton lasing is observed from a bulk GaN-based microcavity diode. The low nonlinear threshold for polariton lasing occurs at 169 A/cm^{2} in the light-current characteristics, accompanied by a collapse of the emission linewidth and small blueshift of the emission peak. Measurement of angle-resolved luminescence, polariton condensation and occupation in momentum space, and output spatial coherence and polarization have also been made. A second threshold, due to conventional photon lasing, is observed at an injection of 44 kA/cm^{2}.
Article
Elastic scattering of exciton polaritons in planar semiconductor microcavities has been used to create X-NOR logic gates. Polaritons with identical linear polarization scatter preferentially at the right angle and rotate their polarization by 90°. On the other hand, scattering of polaritons having orthogonal linear polarizations is suppressed. We show that these effects are a consequence of the multiple scattering in microcavities which involve three and more polaritons. The theory quantitatively reproduces the experimental data of C. Leyder et al. Phys. Rev. Lett. 99 196402 (2007)].