Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Jurnal Tentang Mekanika Statistik

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

ARTICLE

Received 26 Oct 2015 | Accepted 16 Mar 2016 | Published 19 Apr 2016

DOI: 10.1038/ncomms11340

OPEN

Calorimetry of a BoseEinstein-condensed
photon gas
Tobias Damm1, Julian Schmitt1, Qi Liang1,w, David Dung1, Frank Vewinger1, Martin Weitz1 & Jan Klaers1,w

Phase transitions, as the condensation of a gas to a liquid, are often revealed by a


discontinuous behaviour of thermodynamic quantities. For liquid helium, for example, a
divergence of the specic heat signals the transition from the normal uid to the superuid
state. Apart from liquid helium, determining the specic heat of a Bose gas has proven to be a
challenging task, for example, for ultracold atomic Bose gases. Here we examine the
thermodynamic behaviour of a trapped two-dimensional photon gas, a system that allows us
to spectroscopically determine the specic heat and the entropy of a nearly ideal Bose gas
from the classical high temperature to the Bose-condensed quantum regime. The critical
behaviour at the phase transition is clearly revealed by a cusp singularity of the specic heat.
Regarded as a test of quantum statistical mechanics, our results demonstrate a quantitative
agreement with its predictions at the microscopic level.

1 Institut fu
r Angewandte Physik, Atominstitut, Institute of Quantum Electronics, Universitat Bonn, Wegelerstrasse 8, 53115 Bonn, Germany. w Present
addresses: Atominstitut, TU Wien, Stadionallee 2, 1020 Vienna, Austria (Q.L.); Institute of Quantum Electronics, ETH Zurich, Auguste-Piccard-Hof 1, 8093
Zurich, Switzerland (J.K.). Correspondence and requests for materials should be addressed to J.K. (email: jklaers@phys.ethz.ch).

NATURE COMMUNICATIONS | 7:11340 | DOI: 10.1038/ncomms11340 | www.nature.com/naturecommunications

ARTICLE

NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11340

elow 2.2 K, liquid helium shows peculiar hydrodynamic


properties, such as a ow without viscosity, the fountain
effect or the formation of vortices1. This transition from a
normal uid to a superuid has been named l-transition, which
originates from the fact that plotting the heat capacity versus
temperature2 results in a graph resembling the greek letter
l. Soon after this discovery, it has been proposed that superuid
helium forms a macroscopic matter wave as a consequence of
BoseEinstein condensation3, which describes the condensation
of the ideal (interaction-free) Bose gas at low temperatures due to
quantum statistics4. This idea proved to be fruitful despite the fact
that liquid helium is far from a system of interaction-free
particles5. The impressive progress in the cooling of dilute atomic
gases has paved the way to realize weakly interacting Bose gases at
nano-Kelvin temperatures6,7. Here the relation to BoseEinstein
condensation has been immediately clear. Interestingly, in
contrast to liquid helium and a recent measurement of a
strongly interacting atomic Fermi gas8, these systems have not
allowed for detailed calorimetric studies up to now9. Evidence for
a non-classical specic heat has been reported10,11, but the
accuracy obtained in experiments with weakly interacting atomic
Bose gases has not been sufcient for an unambiguous
determination of the temperature dependence of the heat
capacity.
Following ultracold atomic Bose gases, other physical systems
have been demonstrated to undergo BoseEinstein condensation,
for example, gases of excitonpolaritons12,13, magnons14 and, in
previous work of our group, photons1517. In contrast to a threedimensional thermal photon gas as Plancks blackbody radiation,
photons can exhibit BoseEinstein condensation, if the
thermalization process is restricted to two motional degrees of
freedom. Experimentally, this situation has been realized in a
microcavity enclosing a dye medium, designated as a room
temperature heat bath for the photon gas. Detailed experimental
studies of the thermalization16 and condensation process15, as
well as the quantum statistics of the photon condensate17, have
revealed the signatures of an almost ideal Bose gas.
Here we report a measurement of the calorimetric properties of
a BoseEinstein-condensed photon gas, in particular, the
temperature dependence of the specic heat and entropy from
the classical high temperature to the quantum-degenerate regime
at low temperatures. At the phase transition, the observed specic

heat shows a cusp singularity, illustrating critical behaviour for a


photon gas analogous to the l-transition of liquid helium.
Results
Two-dimensional photon gas in a dye microcavity. In our
experiment (Fig. 1a), photons are captured inside a microcavity
consisting of two spherically curved mirrors while repeatedly
being absorbed and re-emitted by the embedded dye medium.
The cavity length is of the same order as the wavelength itself,
which causes a large frequency gap between the longitudinal
resonator modes (free spectral range), comparable to the emission
bandwidth of the dye molecules (Fig. 1b). In this situation, the
resonator becomes populated by photons of a single longitudinal
mode number q only, for example, q 8. While the longitudinal
mode number is frozen out, the photons may populate a multitude of transversally excited cavity modes, for example, the
TEM8xy sub-spectrum, which makes the photon gas effectively
two-dimensional. The photon energy-momentum-relation
acquires a quadratic form, resembling that of a massive particle,
and a trapping potential for the photon gas is induced by the
mirror curvature. One can show that the photon gas conned in
the resonator is formally equivalent to a harmonically trapped
two-dimensional gas of massive bosons18,19, described by the
dispersion
E  mc=n2 2 =2mk2x k2y mO2 =2x2 y2 ; 1
with spatial coordinates x and y, transverse wave vector
components kx and ky, trapping frequency O and an effective
mass m :oc(n/c)2, where :oc is the photon energy in the cavity
ground mode with n as the refractive index of the medium and
c as the vacuum speed of light. Thermal equilibrium of the
photon gas with the cavity environment at room temperature is
achieved via repeated absorption and emission processes by the
dye molecules, which establishes a thermal contact between
photon gas and optical medium18,20,21. Other than in a blackbody
radiator, the thermalization process allows for an independent
adjustment of temperature and photon number, for example, by
(initial) optical pumping, which eventually goes back to a
separation of energy scales of photon energy and thermal
energy. In our experiment, the BoseEinstein condensation is
triggered by increasing the photon number above the saturation

Excitation
Mirror

Cavity
spectrum

Degeneracy
(a.u.)

TEM8xy

c

D0=q /2

Dye

Strength
(a.u.)

Dye
Emission

400

Absorption

500
600
(2THz)

Figure 1 | BoseEinstein condensation of a two-dimensional photon gas. (a) Photons are captured inside a microcavity consisting of two spherically
curved mirrors and get repeatedly absorbed and re-emitted by the embedded dye medium, leading to a thermalization of the photon gas to the temperature
of the resonator (room temperature). (b) The short cavity length causes a large frequency gap between the longitudinal resonator modes (free spectral
range) of order of the emission bandwidth of the dye molecules. In this situation, the resonator becomes populated by photons of a single longitudinal
mode number only, here q 8. However, the photons may still populate a multitude of transversally excited cavity modes (TEM8xy sub-spectrum), which
effectively makes the photon gas two-dimensional. Above a critical photon number, the photon gas undergoes a BoseEinstein condensation, leading to a
massive population of the cavity ground mode (TEM800). Thermodynamic information is obtained by spectroscopically analysing the photon energy
distribution across the phase transition.
2

NATURE COMMUNICATIONS | 7:11340 | DOI: 10.1038/ncomms11340 | www.nature.com/naturecommunications

ARTICLE

NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11340

level at a given temperature. The corresponding critical particle


number is given by2224
2

1,000

100
10

0.75
8

0.5

4
2

n0 /N

which for typical experimental parameters corresponds to


NcE90,000.
One of the benets of the given experimental system is
that one can easily interpolate between equilibrium and
non-equilibrium experimental conditions. Parameters, such as
mirror reectivity, dye concentration, cavity tuning and pump
geometry, can be chosen such that gain and dissipation either
signicantly contribute to the system dynamics or effectively
drop out of it16,20,2527. In this study, we have concentrated on
the equilibrium properties of the system. Details on the
corresponding experimental parameters can be found in the
Methods section. In our experiments, the optical medium is
pumped with a spectrally off-resonant laser source at a
wavelength of lexc 532 nm, having a relatively large beam
diameter of B150 mm to keep the excitation of the medium
nearly spatially homogeneous. Two acousto-optical modulators
(AOMs) are used to chop the pump light to long pulses of
400 ns length with a repetition rate of 400 Hz. The AOMs
further control the intensity of each light pulse that allows us to
adjust the average photon number with respect to the critical
photon number. In our experiment, we quickly ramp the total
photon number from typically 30,000 to 550,000 photons
within 250 ms.
The thermodynamic properties of the system are experimentally accessible by spectroscopic means. To obtain the intracavity spectral distribution of the photons, we measure the
spectral distribution outside the resonator and divide by the
(wavelength-dependent) transmission coefcient of the mirrors.
For this, the emitted cavity light is collimated and coupled to a
monochromator (4f set-up), where it is spectrally decomposed
by a diffraction grating (1,200 grooves per mm), giving an
overall resolution of 0.5 nm. With this set-up, it is possible to
capture the photon gas spectrum in the wavelength region
starting from E550 nm to the cavity cutoff at lc 580 nm
corresponding to an energy range of E4 kBT. The latter
comprises the condensate population and E95% of photons
in excited cavity modes. The E5% most energetic photons of
the thermal cloud are not experimentally resolvable with the
present set-up.
Figure 2a shows spectra obtained at xed temperature
T 300 K for total photon numbers ranging from NE30,000 to
NE550,000 (varying chemical potential). The observed spectra
generally are in good agreement with BoseEinstein distributions,
with residual deviations at the lower photon energy part due to
imperfect reabsorption. The measured energy distributions can be
used to obtain full thermodynamic information of the twodimensional photon gas. All of the derived quantities will be
measured for constant volume, which here means constant
(inverse squared) trapping frequency, and xed absolute
temperature T. As a rst step, we have determined the condensate
fraction n0/N as a function of the reduced temperature T/Tc
(Fig. 2b). From equation (2) follows that the reduced temperature
T/Tc is related to the total particle number by T/Tc (Nc/N)1/2,
with the total particle number N being obtained by integrating
over the spectrum. The experimentally derived condensate
fractions are slightly below the theoretical expectations shown
by the solid line in Fig. 2b, describing an inverse parabola
n0/N 1  (T/Tc)2. This stems from an imperfect saturation of
the population in excited photon modes, which has already
been observed in previous measurements15, and potentially
originates from a weak (thermo-optically induced) photon
self-interaction28.

Signal (a.u.)

Nc  p =3kB T= O ;
2

0.25
0

0.1

0
560
570
580
Wavelength  (nm)

0.5

1.5

Temperature T/Tc

Figure 2 | Spectral photon distribution and condensate fraction.


(a) Distribution of photon energies for increasing total photon number
(circles). For clarity, the spectra have been vertically shifted. The observed
spectra agree well with the expected 300 K BoseEinstein distribution
functions (solid lines). (b) Condensate fraction n0/N versus the reduced
temperature T/Tc along with the theoretical expectation n0/N 1  (T/Tc)2.
Due to an imperfect saturation of the population of the thermal cloud, the
condensate fraction is observed to be systematically below the theoretically
expected values.

Caloric and entropic properties. We next determine the average


energy per photon U/N, with the zero point of the energy scale
being set to the energy :oc hc/lc of the cavity ground state
(TEM00 mode), corresponding to a condensate wavelength of
lc 580 nm. On the basis of the experimentally obtained
spectral photon distribution n(l) in the wavelength regime
lE550580
we extrapolate the total internal energy to be
R 580nm,
nm n(l) hc (l  1  l  1) dl, with h as Planck
UEk  550
c
nm
constant and c as the vacuum speed of light. The extrapolation
factor k is uniquely determined by the assumption that the
spectral distribution continues to be Boltzmann like in the
experimentally not resolved wavelength regime lo550 nm,
containing the E5% most energetic photons of the thermal
cloud. The latter sets a value of k U/U(l4550 nm)E1.19, with
U(l4550 nm), denoting the energy contribution of photons with
wavelength l4550 nm.
In Fig. 3a, the average energy U/N normalized to the
characteristic energy at criticality kBTc is plotted versus
the reduced temperature T/Tc, showing good agreement with
the theoretical expectations for an ideal Bose gas (solid lines). At
higher temperatures T4Tc, the energy shows a linear scaling with
temperature, as expected in the classical limit, where Maxwell
Boltzmann statistics applies. In the vicinity of the condensation
threshold, the energy curve changes slope, as is more clearly
revealed in the graph of the heat capacity in Fig. 3b. The
data points are obtained by numerically differentiating the
measured energy curve with respect to temperature, following
C qU/qT kB q(U/kBTc)/q(T/Tc). The obtained heat capacity
data (circles) shows the characteristic l-like shape as predicted
theoretically (solid lines). In the high-temperature (classical)
regime, C reaches a limiting value of 2kB per photon. The latter
stems from the four degrees of freedom, two kinetic (kx and ky)
and two potential (x and y), which quadratically enter the
photon energy of equation (1) and each contribute with kB/2
to the specic heat (equipartition theorem). At criticality, the
heat capacity shows a cusp with a maximum value of
C(Tc)/N (3.80.3)kB, slightly below the theoretically expected
value in the thermodynamic limit of 6z(3)/z(2)kBE4.38kB
(refs 2224). Most of this discrepancy can be explained by the
nite size effects. From an exact numerical evaluation of the
BoseEinstein distribution function, one can obtain the specic
heat for nite system sizes (solid lines in Fig. 3b). For the given

NATURE COMMUNICATIONS | 7:11340 | DOI: 10.1038/ncomms11340 | www.nature.com/naturecommunications

ARTICLE
a

NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11340

104

0.5

Entropy S/NkB

1.5

Specific heat C/NkB

Pressure P/ 2NkBTc

Energy U/NkBTc

105

3
2

0.5
1
1.5
Temperature T/Tc

0.5

1.5

Temperature T/Tc

0.5

1.5

Temperature T/Tc

Figure 3 | Caloric and entropic properties. (a) Internal energy (per photon) U normalized to the characteristic energy kBTc as a function of the reduced
temperature T/Tc. In the classical high-temperature regime, energy scales linearly with temperature as is expected from MaxwellBoltzmann-like statistics.
The formation of the condensate is signalled by a change in slope close to T Tc. For an ideal two-dimensional trapped Bose gas, which is realized in our
experiment in good approximation, the internal energy of the photon gas can moreover be linked to its pressure P by P (1/2)O2U, see axis on the righthand side and the main text for details (O is the trapping frequency). Experiments are carried out at constant temperature T 300 K and photon numbers
ranging from NE30,000 to NE550,000, whereby an increase of the photon number corresponds to a decrease in the critical temperature Tc Tc(N) of
the system. s.e.m.s are smaller than the point size. (b) Specic heat (per photon) versus the reduced temperature T/Tc, showing a cusp singularity at
criticality (circles). In the high-temperature (classical) regime, the specic heat reaches a limiting value of 2kB per photon. At criticality T Tc, the heat
capacity reaches a maximum value of C(Tc)/N (3.80.3)kB. The solid lines give the specic heat of the two-dimensional harmonically trapped ideal Bose
gas for varying total particle numbers. The error bars indicate s.e.m. (c) Entropy per photon as a function of the reduced temperature T/Tc (circles), along
with the theoretically expected ideal Bose gas behaviour (solid line). The data are derived from a numerical integration of the specic heat curve of b (see
text for details). The entropy monotonically decreases for decreasing temperatures, reaching a minimum value of E0.2kB per photon at the lowest
achieved temperature. s.e.m.s are smaller than the point size.

photon numbers near NE90,000 at criticality, the nite size


effects reduce the specic heat maximum to a value of 4.11kB,
which agrees with the measured value within the experimental
uncertainties. For ToTc, the heat capacity monotonically
decreases for smaller temperatures, dropping below the classical
value of 2kB per photon at T/TcE0.7, and being consistent with
reaching zero at T 0, as would be demanded by the third law of
thermodynamics. Presently, the minimum achieved temperature
is T/TcE0.4, corresponding to condensate fractions of up to 84%,
due to limited available pump power.
The specic heat furthermore gives access to other thermodynamic quantities,R as the entropy, which can be determined by
the integral S(T) 0TC(T0 )/T0 dT0 , see Fig. 3c for the corresponding data. As the specic heat C(T) is not known for temperatures
below T/TcE0.4, there is one free parameter, the constant offset
S(T/TcE0.4) that needs to be set to evaluate the integral and to
match the experimental data (circles) to the theoretical prediction
(solid line). The entropy curve monotonically decreases with
decreasing temperature, reaching a minimum value of E0.2kB
per photon for the lowest obtained temperature. Although we
presently cannot access temperatures closer to zero, the observed
drop-off of the entropy curve towards lower temperatures is in
accordance with the third law of thermodynamics. Note that the
entropy per particle (as a function of chemical potential) also has
been experimentally determined for a trapped two-dimensional
atomic Bose gas, reaching entropies as low as 0.06(1)kB (ref. 29).
Discussion
For a gas of non-interacting bosons, which in good approximation is realized in our experiment as demonstrated by the results
given in Figs 2 and 3, one can readily link the internal energy of
the gas to its pressure. In general, in the presence of a trapping
potential, the pressure of a gas becomes position dependent and
thus cannot serve as a global thermodynamic variable. To account
for this, it has been proposed to use two global conjugated
4

variables, harmonic volume V and harmonic pressure P,


respectively11,30. For a harmonically trapped two-dimensional
gas, the harmonic volume is dened as V O  2, with O as the
trapping frequency. This quantity does not have the physical
units of a volume, however, it shares the same scaling with the
trapping geometry as the true volume of the conned gas. The
harmonic pressure then follows via the usual thermodynamic
relation P  qfG/qV, where fG is the grand potential. For a
two-dimensional trapped gas of non-interacting bosons, the
pressure can be shown to be related to the internal energy via
P (1/2)O2U. Thus, Fig. 3a here does not only describe the
energy as a function of temperature U U(T/Tc) (caloric
equation of state) but also delivers the pressure dependence
P P(T/Tc) (thermal equation of state) for a given harmonic
volume V or trapping frequency O.
To conclude, we have determined calorimetric properties of a
BoseEinstein-condensed photon gas, in particular the temperature dependence of energy, heat capacity and entropy. Critical
behaviour of the photon gas is clearly demonstrated by a cusp in
the specic heat curve at the condensation threshold. For the
chosen experimental conditions, for example, sufcient photon
reabsorption and nearly homogenous pump geometry, we do not
observe signicant deviations from the theoretically expected
behaviour of a fully equilibrated Bose gas. In comparison
with other systems exhibiting BoseEinstein condensation, the
here-investigated photon gas comes closest to an ideal, that is,
interaction-free, gas of bosons, allowing to match experimental
results with precise, and even exact, theoretical predictions. In
particular, we nd that the experimentally determined specic
heat of the photon gas agrees with the quantum statistical
predictions down to the level of nite size effects. For the future,
the spectroscopic calorimetry of a quantum-degenerate photon
gas could lead to new experimental schemes for precision
measurements of thermodynamic quantities as the Boltzmann
constant31,32.

NATURE COMMUNICATIONS | 7:11340 | DOI: 10.1038/ncomms11340 | www.nature.com/naturecommunications

ARTICLE

NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11340

Methods
Microcavity set-up. In our experiment, photons are stored inside a microcavity
build up from two gyro-quality mirrors. The dielectric, spherically shaped mirrors
have a reectivity in excess of rE0.99997 in the relevant wavelength region of this
experiment (l 530590 nm), providing a cavity nesse of order of FE105 for the
empty cavity. Both the mirrors share the same radius of curvature of R 1 m and
are typically separated by D0E1.7 mm, corresponding to a free spectral range of
order DlFSRE100 nm, which is comparable to the spectral width of the dye
emission. In this situation, the dye emission is restricted to cavity modes with the
longitudinal wave number q 8, effectively reducing the thermalization dynamics
of the photon gas to the remaining two transversal mode numbers. The effective
mass and trapping frequency introduced in equation (1) depend on the cavity
geometry and typically take values of mE6.7  10  36 kg and OE2p  36.5 GHz in
our experiment. The cavity geometry is stabilized passively, by mechanical contact
of the two mirrors that strongly damps fast mechanical oscillations, as well as
actively utilizing a piezo translation stage that counteracts long time drifts of the
resonance.
As a heat bath for the photon gas, we use a ltered solution of 10  3 mol l  1
rhodamine 6G ethylene glycol (uorescence quantum yield ZE0.95, index of
refraction of the solvent n 1.43). This dye fulls the KennardStepanov law
B21(o)/B12(o)Eexp(  (o  oZPL)/kBT), relating the Einstein coefcients of
absorption B12(o) and emission B21(o) at a given frequency o to the Boltzmann
factor of that frequency (oZPL is the zero-phonon line of the dye), which is essential
for the lightmatter thermalization process. The optical medium is spatially
homogeneously pumped by a spectrally off-resonant laser system at a wavelength
of lexc 532 nm under an angle of E45 with respect to the optical axis exploiting
the rst reectivity minimum at higher angles of incidence. To avoid excess
population of long-lived dye triplet states and photobleaching two AOM are used
to chop the pump light to pulses of 400 ns length with a repetition rate of 400 Hz.

References
1. Tilley, D. R. & Tilley, J. Superuidity and Superconductivity (CRC Press, 1990).
2. Lipa, J. A., Nissen, J. A., Stricker, D. A., Swanson, D. R. & Chui, T. C. P. Specic
heat of liquid helium in zero gravity very near the lambda point. Phys. Rev. B
68, 174518 (2003).
3. London, F. The l-phenomenon of liquid helium and the Bose-Einstein
degeneracy. Nature 141, 643644 (1938).
4. Einstein, A. Quantum theory of the monoatomic ideal gas, part II. Sitzber.
Preuss. Akad. 1, 314 (1925).
5. Penrose, O. & Onsager, L. Bose-Einstein condensation and liquid helium. Phys.
Rev. 104, 576584 (1956).
6. Anderson, M. H., Ensher, J. R., Matthews, M. R., Wieman, C. E. & Cornell, E. A.
Observation of Bose-Einstein condensation in a dilute atomic vapor. Science 269,
198201 (1995).
7. Davis, K. B. et al. Bose-Einstein condensation in a gas of sodium atoms. Phys.
Rev. Lett. 75, 39693973 (1995).
8. Ku, M. J. H., Sommer, A. T., Cheuk, L. W. & Zwierlein, M. Revealing the
superuid lambda transition in the universal thermodynamics of a unitary
Fermi gas. Science 335, 563567 (2012).
9. Blakie, P. B., Toth, E. & Davis, M. J. Calorimetry of Bose-Einstein condensates.
J. Phys. B: At. Mol. Opt. Phys. 40, 32733282 (2007).
10. Ensher, J. R., Jin, D. S., Matthews, M. R., Wieman, C. E. & Cornell, E. A.
Bose-Einstein condensation in a dilute gas: measurement of energy and
ground-state occupation. Phys. Rev. Lett. 77, 49844987 (1996).
11. Shiozaki, R. F. et al. Measuring the heat capacity in a Bose-Einstein
condensation using global variables. Phys. Rev. A 90, 043640 (2014).
12. Kasprzak, J. et al. Bose-Einstein condensation of exciton polaritons. Nature
443, 409414 (2006).
13. Balili, R., Hartwell, V., Snoke, D., Pfeiffer, L. & West, K. Bose-Einstein
condensation of microcavity polaritons in a trap. Science 316, 10071010
(2007).
14. Demokritov, S. O. et al. Bose-Einstein condensation of quasi-equilibrium
magnons at room temperature under pumping. Nature 443, 430433 (2006).
15. Klaers, J., Schmitt, J., Vewinger, F. & Weitz, M. Bose-Einstein condensation of
photons in an optical microcavity. Nature 468, 545548 (2010).

16. Klaers, J., Vewinger, F. & Weitz, M. Thermalization of a two-dimensional


photonic gas in a white wall photon box. Nat. Phys. 6, 512515 (2010).
17. Schmitt, J. et al. Observation of grand-canonical number statistics in a photon
Bose-Einstein condensate. Phys. Rev. Lett. 112, 030401 (2014).
18. Klaers, J., Schmitt, J., Damm, T., Vewinger, F. & Weitz, M. Statistical physics of
Bose-Einstein-condensed light in a dye microcavity. Phys. Rev. Lett. 108,
160403 (2012).
19. Nyman, R. A. & Szymanska, M. H. Interactions in dye-microcavity photon
condensates and the prospects for their observation. Phys. Rev. A 89, 033844
(2014).
20. Kirton, P. & Keeling, J. Nonequilibrium model of photon condensation. Phys.
Rev. Lett. 111, 100404 (2013).
21. de Leeuw, A.-W., Stoof, H. T. C. & Duine, R. A. Schwinger-Keldysh theory for
Bose-Einstein condensation of photons in a dye-lled optical microcavity. Phys.
Rev. A 88, 033829 (2013).
22. Haugset, T., Haugerud, H. & Andersen, J. O. Bose-Einstein condensation in
anisotropic harmonic traps. Phys. Rev. A 55, 29222929 (1997).
23. Klunder, B. & Pelster, A. Systematic semiclassical expansion for harmonically
trapped ideal Bose gases. Eur. Phys. J. B 68, 457465 (2009).
24. Kozhevnikov, A. A. Bose gas in power-like spherically symmetric potential in
arbitrary spatial dimensionality. Acta Phys. Pol. B 43, 20892102 (2012).
25. Marelic, J. & Nyman, R. A. Experimental evidence for inhomogeneous
pumping and energy-dependent effects in photon Bose-Einstein condensation.
Phys. Rev. A 91, 033813 (2015).
26. Kirton, P. & Keeling, J. Thermalization and breakdown of thermalization in
photon condensates. Phys. Rev. B 91, 033826 (2015).
27. Keeling, J. & Kirton, P. Spatial dynamics, thermalization, and gain clamping in
a photon condensate. Phys. Rev. A 93, 013829 (2016).
28. Tammuz, N. et al. Can a Bose gas be saturated? Phys. Rev. Lett. 106, 230401
(2011).
29. Yefsah, T., Desbuquois, R., Chomaz, L., Gunter, K. J. & Dalibard, J. Exploring
the thermodynamics of a two-dimensional Bose gas. Phys. Rev. Lett. 107,
130401 (2011).
30. Romero-Rochin, V. Equation of state of an interacting Bose gas conned by a
harmonic trap: the role of the harmonic pressure. Phys. Rev. Lett. 94, 130601
(2005).
31. de Podesta, M. et al. A low-uncertainty measurement of the Boltzmann
constant. Metrologia 50, 354376 (2013).
32. Castrillo, A. et al. The Boltzmann constant from the shape of a molecular
spectral line. J. Mol. Spectrosc. 300, 131138 (2014).

Acknowledgements
We thank James Anglin and Henk Stoof for valuable discussions. This work has been
nancially supported by the DFG (We1748-17) and the ERC (INPEC).

Author contributions
All authors contributed extensively to the ndings presented in this paper.

Additional information
Competing nancial interests: The authors declare no competing nancial interests.
Reprints and permission information is available online at http://npg.nature.com/
reprintsandpermissions/
How to cite this article: Damm, T. et al. Calorimetry of a BoseEinstein-condensed
photon gas. Nat. Commun. 7:11340 doi: 10.1038/ncomms11340 (2016).
This work is licensed under a Creative Commons Attribution 4.0
International License. The images or other third party material in this
article are included in the articles Creative Commons license, unless indicated otherwise
in the credit line; if the material is not included under the Creative Commons license,
users will need to obtain permission from the license holder to reproduce the material.
To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/

NATURE COMMUNICATIONS | 7:11340 | DOI: 10.1038/ncomms11340 | www.nature.com/naturecommunications

You might also like