arxiv: v1 [cond-mat.quant-gas] 10 Jan 2011

Similar documents
Hexatic and microemulsion phases in a 2D quantum Coulomb gas

Quantum Monte Carlo Simulations of Exciton Condensates

Wigner crystallization in mesoscopic semiconductor structures

arxiv:cond-mat/ v2 7 Dec 1999

Bound states of two particles confined to parallel two-dimensional layers and interacting via dipole-dipole or dipole-charge laws

Polariton Condensation

Structure Formation in Strongly Correlated Coulomb Systems

Microcavity Exciton-Polariton

Quantum Properties of Two-dimensional Helium Systems

arxiv: v3 [cond-mat.quant-gas] 5 May 2015

Spontaneous Symmetry Breaking in Bose-Einstein Condensates

Phase transitions in Bi-layer quantum Hall systems

Strongly correlated Cooper pair insulators and superfluids

Clas s ical and path integral Monte Carlo s imulation of charged particles in traps

Landau s Fermi Liquid Theory

Topological defects and its role in the phase transition of a dense defect system

Observing Wigner Crystals in Double Sheet Graphene Systems in Quantum Hall Regime

Intensity / a.u. 2 theta / deg. MAPbI 3. 1:1 MaPbI 3-x. Cl x 3:1. Supplementary figures

Phase Transitions in Relaxor Ferroelectrics

lattice that you cannot do with graphene! or... Antonio H. Castro Neto

Spectroscopy of. Semiconductors. Luminescence OXFORD IVAN PELANT. Academy ofsciences of the Czech Republic, Prague JAN VALENTA

arxiv: v1 [cond-mat.dis-nn] 31 Aug 2011

ELECTRIC FIELD EFFECTS ON THE EXCITON BOUND TO AN IONIZED DONOR IN PARABOLIC QUANTUM WELLS

arxiv: v1 [cond-mat.stat-mech] 1 Oct 2009

Entropic Crystal-Crystal Transitions of Brownian Squares K. Zhao, R. Bruinsma, and T.G. Mason

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013

Preface Introduction to the electron liquid

14. Structure of Nuclei

Physics 127c: Statistical Mechanics. Application of Path Integrals to Superfluidity in He 4

Physics 127a: Class Notes

1 Superfluidity and Bose Einstein Condensate

LOCAL MOMENTS NEAR THE METAL-INSULATOR TRANSITION

Quantum spin systems - models and computational methods

Deconfined Quantum Critical Points

6 Kosterlitz Thouless transition in 4 He thin films

Vortices and vortex states of Rashba spin-orbit coupled condensates

Infinitely long-range nonlocal potentials and the Bose-Einstein supersolid phase

Design and realization of exotic quantum phases in atomic gases

Supersolidity of excitons

Strongly correlated systems in atomic and condensed matter physics. Lecture notes for Physics 284 by Eugene Demler Harvard University

Collective Effects. Equilibrium and Nonequilibrium Physics

Cold Polar Molecules and their Applications for Quantum Information H.P. Büchler

PHYS3113, 3d year Statistical Mechanics Tutorial problems. Tutorial 1, Microcanonical, Canonical and Grand Canonical Distributions

High-Temperature Criticality in Strongly Constrained Quantum Systems

Two-Dimensional Spin-Polarized Hydrogen at Zero

5 Topological defects and textures in ordered media

Supplementary Figure S1 Definition of the wave vector components: Parallel and perpendicular wave vector of the exciton and of the emitted photons.

2D Bose and Non-Fermi Liquid Metals

Tunneling Spectroscopy of Disordered Two-Dimensional Electron Gas in the Quantum Hall Regime

Degeneracy Breaking in Some Frustrated Magnets. Bangalore Mott Conference, July 2006

Topological Phase Transitions

Superfluidity and Condensation

Splitting of a Cooper pair by a pair of Majorana bound states

The Hubbard model in cold atoms and in the high-tc cuprates

Phase transitions and critical phenomena

Chapter 6. Phase transitions. 6.1 Concept of phase

766 Liu Bin et al Vol. 12 melting transition of the plasma crystal. Experimental investigations showed the melting transition from a solid-state struc

The phases of matter familiar for us from everyday life are: solid, liquid, gas and plasma (e.f. flames of fire). There are, however, many other

SPINTRONICS. Waltraud Buchenberg. Faculty of Physics Albert-Ludwigs-University Freiburg

Atomic Structure and Processes

Quantised Vortices in an Exciton- Polariton Condensate

arxiv:cond-mat/ v3 [cond-mat.supr-con] 28 Oct 2007

Optical Properties of Lattice Vibrations

First-Principles Calculation of Exchange Interactions

Strongly Correlated Physics With Ultra-Cold Atoms

Low dimensional quantum gases, rotation and vortices

The fate of the Wigner crystal in solids part II: low dimensional materials. S. Fratini LEPES-CNRS, Grenoble. Outline

Dynamical Condensation of ExcitonPolaritons

Examples of Lifshitz topological transition in interacting fermionic systems

Superfluidity in bosonic systems

Strongly Correlated 2D Quantum Phases with Cold Polar Molecules: Controlling the Shape of the Interaction Potential

In-class exercises. Day 1

Bose-Einstein Condensate: A New state of matter

Minimal Update of Solid State Physics

ON QUARK ATOMS, MOLECULES AND CRYSTALS * 3. Department of Nuclear Physics. Weizmann Institute of Science. Abstract

Structure and stability of trapped atomic boson-fermion mixtures

UNCLASSIFIED UNCLASSIFIED

ECE440 Nanoelectronics. Lecture 07 Atomic Orbitals

Self-compensating incorporation of Mn in Ga 1 x Mn x As

David Snoke Department of Physics and Astronomy, University of Pittsburgh

doi: /PhysRevLett

Ginzburg-Landau theory of supercondutivity

Spontaneous Spin Polarization in Quantum Wires

Simulation of two-dimensional melting of Lennard-Jones solid

Spin liquids in frustrated magnets

STIMULATED RAMAN ATOM-MOLECULE CONVERSION IN A BOSE-EINSTEIN CONDENSATE. Chisinau, Republic of Moldova. (Received 15 December 2006) 1.

Microscopic Deterministic Dynamics and Persistence Exponent arxiv:cond-mat/ v1 [cond-mat.stat-mech] 22 Sep 1999

Phases of Na x CoO 2

Supplementary Information for Observation of dynamic atom-atom correlation in liquid helium in real space

Ultracold Fermi and Bose Gases and Spinless Bose Charged Sound Particles

221B Lecture Notes Spontaneous Symmetry Breaking

Giant Enhancement of Quantum Decoherence by Frustrated Environments

arxiv:cond-mat/ v1 [cond-mat.mes-hall] 26 Sep 2001

Trapping Centers at the Superfluid-Mott-Insulator Criticality: Transition between Charge-quantized States

Clusters and Percolation

Supplementary Figure 1 Level structure of a doubly charged QDM (a) PL bias map acquired under 90 nw non-resonant excitation at 860 nm.

WORLD SCIENTIFIC (2014)

Luigi Paolasini

arxiv: v1 [cond-mat.str-el] 24 Oct 2011

2D Electrostatics and the Density of Quantum Fluids

Transcription:

Crystallization of an exciton superfluid J. öning, 1 A. Filinov, 1 and M. onitz 1, 1 Institut für Theoretische Physik und Astrophysik, Christian-Albrechts-Universität, Leibnizstr. 15, D-24098 Kiel, Germany arxiv:1101.1757v1 [cond-mat.quant-gas] 10 Jan 2011 (Dated: January 11, 2011) Abstract Indirect excitons pairs of electrons and holes spatially separated in semiconductor bilayers or quantum wells are known to undergo ose-einstein condensation and to form a quantum fluid. Here we show that this superfluid may crystallize upon compression. However, further compression results in quantum melting back to a superfluid. This unusual behavior is explained by the effective interaction potential between indirect excitons which strongly deviates from a dipole potential at small distances due to many-particle and quantum effects. ased on first principle path integral Monte Carlo simulations, we compute the complete phase diagram of this system and predict the relevant parameters necessary to experimentally observe exciton crystallization in semiconductor quantum wells. 1

Quantum coherence of bosonic particles is one of the most striking macroscopic manifestations of the laws of quantum mechanics governing the microworld. The discovery of ose-einstein condensation in atomic vapors [1] was followed by the observation of condensation of bosonic quasiparticles in condensed matter excitons. Here we mention early claims for three-dimensional (3D) semiconductors [2], electron bilayers in a quantizing magnetic field, e.g. [3, 4], excitonpolaritons in microcavities [5, 6] and so-called indirect excitons formed from spatially separated electrons and holes [7 11]. Not only the bosonic gas phase was observed but also the formation of a quantum ose liquid an exciton superfluid with its peculiar loss of friction could recently be verified [4, 6]. Thus it is tempting to ask whether there exists also a solid phase of bosons. The key properties of a crystal are particle localization and long-range spatial ordering. To achieve spontaneous crystallization requires to find a ose system with sufficiently strong and long range pair interaction (here we do not consider particle localization induced by an external field in an optical lattice or cavity [12, 13]). However, the vast majority of previous experimental investigations have been performed in the regime of weak nonideality, where the interaction energy is small compared to the quantum kinetic energy. Therefore, promising candidates for a bosonic solid are atoms or molecules with dipole interaction [14] or excitons. Here, indirect excitons offer a number of attractive features: a strong dipole-type interaction, the suppression of biexciton or trion formation, the comparatively long radiative life time (on the order of microseconds) and the external controllability of the density and dipole moment via an electric field perpendicular to the quantum well plane, e.g. [10, 11, 15]. In this paper we present clear evidence for the existence of a crystal of indirect excitons in semiconductor quantum wells. We compute its full phase diagram and reveal the parameters for its experimental verification. Our predictions are based on first principle path integral Monte Carlo (PIMC) simulations. ut in contrast to previous quantum Monte Carlo studies which predicted crystallization in model systems such as electron-hole bilayers, e.g. [8, 16], or two-dimensional dipole systems, e.g. [17, 18], here we use realistic parameters typical for indirect excitons. In particular, we fully take into account the finite quantum well width, the composite character of the excitons and the different masses of electrons and holes. This turns out to be of crucial importance for the exciton-exciton interaction which strongly departs from a dipole potential at small distances. As a direct consequence we observe that the exciton crystal exists only in a finite density interval and undergoes quantum melting both at high and low density. Furthermore when the exciton superfluid crystallizes to form a solid, quantum coherence is lost abruptly, i.e. there is 2

no supersolid exciton phase. MODEL We consider a semiconductor quantum well (QW) of width L containing N e = N h electrons and holes in the conduction and valence band, respectively, which are created by an optical pulse [19]. Application of an electrostatic field of strength E perpendicular to the QW plane created e.g. by a tip electrode allows to spatially separate electrons and holes to different edges of the QW. y varying E this separation can be changed between 0 and L giving rise to a variable dipole moment d. At the same time, the field also provides lateral confinement and a variable particle density, via the quantum confined Stark effect, for details of the setup, see [15]. Finally, the system is kept in thermal equilibrium at a finite temperature T which does not exceed a few percent of the binding energy of an electron-hole pair, thus all electrons and holes will be bound in N X = N e indirect excitons [20]. The thermodynamic properties of this system are fully described by the density operator of N e electrons and N h holes, ˆρ A N e,n h = A{e βĥ/z}, where Z is the partition function, β = 1/(k T ) and A denotes full anti-symmetrization among all electronic and hole variables. Ĥ is the full Hamiltonian containing kinetic energy, the interaction with the external electric field and all Coulomb pair interactions between the 2N e charged particles. Under the present conditions of strongly bound indirect excitons with parallel dipole moments resulting in a strong exciton-exciton repulsion the very complicated evaluation of the density operator ˆρ A N e,n h can be substantially simplified. As was shown in [22] the system can be mapped onto N X excitons which can be treated as composite spin polarized bosons [23] where deviations from the ose statistics (arising from the original Fermi statistics of electrons and holes) have been found negligible. Thus, the density operator is reduced to a fully symmetrized one of N X excitons, ˆρ S N X. Furthermore, all pair interactions can be properly averaged along the QW width giving rise to an effective (d-dependent) exciton-exciton interaction V XX. As a result the system Hamiltonian entering ˆρ S N X becomes Ĥ eff = 2 2M X N X i=1 2 r i + i<j V XX (r ij ; d), (1) where M X = m e + m h is the in-plane effective mass, r i the center of mass (com) coordinate of the i th exciton and r ij = r i r j denotes the com distance between two excitons. 3

E [Ha ] a 0.6 0.5 0.4 0.3 0.2 0.1 0 V XX E S XX = E e + 1/r hh E A XX E XX = E e +V hh PIMC: E XX E e (m h /m e = 2.46) E e (m h ) Dipole b c r hh = 3 5 10 15 20 35 r hh = 3 15 35 n e 0.4 0.3 0.2 0.1 0 g ee 0.4 0.3 0.2 0.1 0.1 0 5 10 15 20 25 distance r in a 0 50 distance r in a 0 Figure 1: Exciton interaction potential V XX for a dipole moment d = 13.3a. (a) V XX (green dashed line) is compared to the exciton interaction energy E XX in various approximations (see Methods section): results of our model (PIMC), results computed according to Ref. [24] for symmetric (EXX S ) and antisymmetric (E A XX ) electronic states assuming m h and improved results for a realistic mass ratio m h /m e = 2.46 (E XX ). Also shown is the electron contribution E e to the exciton interaction energy and the dipole potential d 2 /r 3. Two vertical lines indicate the boundaries of the exciton crystal. (b) Radial electron density n e (r), relative to the mid-point of the two holes at ± 1 2 r hh) for several holes separations r hh. (c) Electron pair distribution function g ee (r ee ) for the values r hh in (b). We have carefully verified this model by comparing the potential V XX with results of Schindler and Zimmermann [24] which we improved by taking into account a finite hole mass. The results are summarized in Fig. 1 and in the Methods section and confirm the accuracy in the whole density range considered in the present work. Figure 1 shows V XX for a typical experimentally feasible e-h separation d = 13.3a. At large distances, r d, V XX practically coincides with the classical dipole potential, V D = d 2 /r 3, so we expect the system to behave like 2D polarized dipoles, at low densities. At smaller distances, r < d, however, V XX is more akin to a Coulomb potential which is due to the hole repulsion, cf. red dashed curve in Fig. 1. Finally, due to quantum effects arising from the finite hole mass, for r d, V XX approaches a finite value at zero distance. 4

NUMERICAL RESULTS Using PIMC simulations with ˆρ s N X and the Hamiltonian (1) the thermodynamic properties of the N X strongly correlated excitons can be efficiently computed with full account of all interactions, quantum and spin effects, without further approximations. elow we will use atomic units, i.e. lengths will be given in units of the electron ohr radius, a = 2 ɛ/(e 2 m e), and energies in units of the electron Hartree, Ha = e 2 /(ɛa ). Of central importance for the crystallization is the coupling (nonideality) parameter, i.e. the ratio of interaction energy to kinetic energy. For a quantum system with Coulomb (dipole) interaction it is given by the rueckner parameter r s (the dipole coupling parameter D), r s r n 1/2, D M X 1 d 2 n 1/2, (2) a m e πrs a 2 where r is the mean interparticle distance and n the exciton density. Note the opposite scaling of r s and D with density. We have performed extensive canonical and grandcanonical 2D PIMC simulations with periodic boundary conditions using N X = 60... 500 excitons. To map out the phase diagram we scan a broad parameter range spanning three orders of magnitude of density and temperature. We first obtain the phase diagram for a fixed value of the dipole moment, corresponding to d = 13.3 a, and after that analyze how the crystal phase boundary changes when d is varied. To detect crystallization we compute the exciton pair distribution function [PDF], g(r). This function equals unity everywhere, in an ideal gas, whereas in the fluid and crystal phase it exhibits increasingly strong modulations signaling spatial localization. Typical examples of g(r) are displayed in the top rows of Figs. 2 and 3 and show clear evidence of exciton localization. The existence of translational long range order (LRO) is detected from the asymptotic behavior of the angle-averaged function g(r) for large r = r. In 2D a possible freezing scenario is given by the Kosterlitz-Thouless-Nelson-Halperin-Young (KTNHY) theory (for an overview see [25]), predicting an exponential (algebraic) decay of the peak heights of g(r) above (below) the melting temperature. Indeed, our simulations find some support for this scenario, see bottom left part of Fig. 2 and fig. S2 of Supplementary material [26]. Similarly, existence of angular hexagonal LRO follows from the asymptotic behavior of the bond angular correlation function, g 6 (r) = ψ6(r)ψ 6 (0), with ψ 6 (r k ) = n 1 nl l l=1 ei6θ kl, where nl is the number of nearest neighbors of a particle located at r k, and Θ kl is their angular distance. We observe a change from an exponential asymptotic of g 6 to a constant which is the expected behavior for a liquid-solid transition, see bottom right part of 5

r/r s (a) (b) (c) g( r) 1 g(r) 1 10 3 10 4 10 3 k T/Ha =1.00 1.05 1.25 1.52 0 g 6 (r) 10 1 10 2 10 5 10 3 6 8 10 15 6 8 10 15 pair distance r/r s Figure 2: Constant density freezing. Top row: 2D PDF g(r ij ) [relative to a fixed particle in the center] for na 2 = 0.0035 and temperatures k T/Ha of 1.74 10 3 (a), 1.38 10 3 (b) and 1.08 10 3 (c). ottom: Radial distribution function g(r) 1 (left) and bond angular order distribution function g 6 (r) (right) at na 2 plot. = 0.0022 for N X = 500. Lines are guide to the eye to visualize an algebraic decay in this log-log Fig. 2. There are some indications for the existence of an hexatic phase coexistence of angular quasi-lro (algebraic decay) and missing translational LRO in a narrow temperature interval, see curves for k T = 1.05 10 3 Ha and k T = 1.25 10 3 Ha. After analyzing emergence of spatial ordering let us now turn to the quantum coherence properties of nonideal indirect excitons. In a 2D ose system cooling leads to sudden emergence of coherence in the liquid phase the normal fluid superfluid transition. The phase boundary is governed by the erezinskii-kosterlitz-thouless (KT) scenario [1] and is given by the condition χ = 4/γ s, for the exciton quantum degeneracy parameter χ nλ 2 where n s k T KT (n s ) = π 2 n s m e M X Ha, (3) = γ s n is the exciton superfluid density. Therefore, a key quantity is the superfluid fraction γ s, where 0 γ s 1. In PIMC simulations, it is directly computed from the statistics of the winding number W [28]: γ s = M X N X 2 β W 2, W = N X β i=1 0 dt dr i(t) dt. (4) 6

r/r s (a) (b) (c) g( r) 1.6 1.4 1.2 1 5 0 (d) (e) (f) 0.8 0.6 0.4 5 γ s 1 0.5 k T 1 = 0.001 Ha k T 2 = 0.0001 Ha 0.2 0 0 0.0003 0.001 0.003 density n [(a ) 2 ] Figure 3: Isothermal freezing and melting of indirect excitons. a)-f): 2D PDF g(r) for k T 1 = 0.001 Ha at densities na 2 of 0.84 10 3 (a), 1.3 10 3 (b), 1.7 10 3 (c), 3.2 10 3 (d), 3.6 10 3 (e), and 4.0 10 3 (f). ottom panel: Superfluid fraction γ s, Eq. (4), vs. density for two temperatures. Symbols are PIMC results, lines are a guide to the eye. The increase of γ s at high density extends over a small finite range of solid-liquid coexistence which is due to the finite particle number in the simulations. Typical simulation results for γ s are shown in the bottom part of Fig. 3. PHASE DIAGRAM We now summarize our findings in the complete phase diagram of indirect excitons in the density temperature plane which is presented in Fig. 4. The phase boundaries are those of a macroscopic system and are obtained by applying a finite size scaling analysis [18], see fig. S1 of Supplementary material. The degeneracy line χ = 1 separates the regions of classical (above the line) and quantum behavior (below). While classical excitons exist only in a fluid (or gas) phase the quantum region is composed of three different phases: a normal fluid, a superfluid and a crystal phase [20]. Correspondingly, there exist two triple points, at the upper left (right) edge of the crystal phase. At high temperature the excitons are in the fluid phase. Cooling leads either into the superfluid or crystal phase. There is no cooling transition from the superfluid to the crystal. At low densities cooling always leads into the superfluid phase; the transition is accompanied by a sudden increase of γ s from zero to a finite value. The phase boundary is substantially below 7

T [Ha ] 0.01 0.001 0.0001 χ = 1 χ = 4 solid r s = 9.4 ± 0.3 superfluid normalfluid T dip T KT D = 17 ± 1 e-h plasma 0.0001 0.001 0.01 density n [(a ) 2 ] Figure 4: Phase diagram of 2D indirect excitons with d = 13.3 a. Circles and squares mark our PIMC results, data for triangles are from [18]. Vertical dashed lines (D = 17 ± 1 and r s = 9.4 ± 0.3) indicate the two density induced quantum freezing (melting) transitions. Filled symbols mark the two triple points. The normal fluid superfluid phase boundary is marked by the red line and is below the ideal estimate T KT according to Eq. (3), cf. thick solid line labeled χ = 4. Line T dip marks the freezing transition of a classical 2D dipole system. The e-h plasma phase is beyond the present analysis. the upper limit T KT (n s = n), Eq. (3), and is in full agreement with our previous analysis for 2D dipoles [18] indicating that the exciton interaction is close to a dipole potential. The picture suddenly changes when the density exceeds na 2 0.00078: the superfluid transition vanishes and, instead, a strong modulation of the PDF is observed signaling crystallization, cf. top row of Fig. 2. The critical density corresponds to a dipole coupling parameter D c = 17 ± 1 which agrees with studies of pure 2D dipole systems [17, 29]. Note that the freezing temperature changes nonmonotonically exhibiting a maximum value T max around na 2 0.002. The superfluid-solid transition is verified by simulating compression along several isotherms. At low temperature and low density, the superfluid fraction γ s starts from a high value until it suddenly drops to zero at the critical density na 2 0.00078, cf. bottom part of Fig. 3. This behavior persists up to zero temperature, cf. Fig. 4. Vanishing of quantum coherence upon crystallization is a general feature and indicates that there is no supersolid phase of indirect excitons. If the temperature is above the left triple point the superfluid fraction is exactly zero, and compression leads to a phase transition from the normal fluid to the crystal phase, cf. γ s for T = 0.001 Ha and the change of the PDF in Fig. 3a c. Interestingly, if the density is increased further, the exciton crystal melts, cf. Fig. 3e,f, this 8

time accompanied by a jump of the superfluid fraction from zero to about 0.9. This indicates isothermal quantum melting to a (partially) superfluid exciton liquid. This occurs at a density of n C a 2 = 0.0036 ± 0.0003 corresponding to r c s = 9.4 ± 0.3 and, again, persist to zero temperature. At temperatures above the right triple point, k T 0.001 Ha, melting and onset of superfluidity are decoupled: first the crystal melts into a normal fluid which becomes a superfluid only at a higher density, cf. Fig. 4. Thus the most striking feature of the exciton phase diagram is the existence of two quantum freezing (melting) transitions, even in the ground state. At low-density excitons undergo pressure crystallization which is characteristic for the behavior of dipole systems or, more generally, for neutral matter composed of atoms or molecules. In addition, at higher densities, there is a second transition: quantum melting by compression. While such an effect is absent in conventional neutral matter it is ubiquitous in Coulomb systems, including the Wigner crystal of the strongly correlated electron gas, ion crystals in the core of white dwarf stars and nuclear matter in the crust of neutron stars. The existence of this quantum melting transition in indirect excitons is due to the peculiar shape of the effective potential V XX : one readily confirms in Fig. 1 that at the critical density where the mean exciton-exciton distance equals 9.4 a, V XX essentially follows the Coulomb repulsion of the holes (red dashed curve). EXPERIMENTAL REALIZATION The results shown in the above figures were computed for d = 13.3 a. Using values from [22], this dipole moment can be achieved, e.g., in a ZnSe quantum well of width L 50 nm or a GaAs quantum well with L 148 nm, both at an electric field strength of E = 20 kv/cm. The density interval for the exciton crystal is estimated as 1.3 10 9 cm 2... 3.6 10 9 cm 2 for GaAs and 8.2 10 9 cm 2... 3.8 10 10 cm 2 for ZnSe. An estimate for the maximum temperature where the crystal can exist is obtained from the classical dipole melting curve, k T dip = c d2 a 2 where c 0.09 [30], and the critical density n C a 2 ( ) na 2 3/2 Ha, (5) = 0.0036 is being used. Taking into account that this value is approximately a factor 2 too high, cf. Fig. 4, we obtain the estimates k T max = 0.17 K (GaAs) and k T max = 0.78 K (ZnSe). These parameters are well within reach of current experiments. A particular advantage is that the upper density limit for exciton crystallization is a 9

d [a ] 20 15 10 5 temperature k T [Ha ] 10 3 10 2 10 1 D c = 17 k T max dip exciton solid r c s = 9.4 d = 13.3a d c = 9.1a 10 4 10 3 10 2 density n [(a ) 2 ] Figure 5: oundaries of the exciton crystal for different dipole moments d. Lower abscissa: density range given by D c = 17 and rs c = 9.4. Upper abscissa: maximum temperature estimated from Tdip max. No solid phase exists for d d c 9.1 a. factor 16 higher than the threshold for an electron Wigner crystal (r s 37). A suitable diagnostics for the crystal can be ragg scattering [15]. Finally, we analyze the dependence of the phase diagram on the dipole moment d which can be varied in a broad range by properly choosing the QW width and the electric field strength. As shown in Fig. 5, an increase of d reduces the lower density limit of the crystal phase whereas the upper boundary remains unchanged. Thus, the crystal phase expands with d, the maximum temperature T max grows quadratically, cf. Eq. (5) and Fig. 5. Finally, there exists a minimum value d c = 9.1 a where the two limiting densities converge, and the crystal phase vanishes. CONCLUSION AND OUTLOOK We have shown that a bosonic many-particle system possesses, besides its weakly nonideal ose condensed gas and its superfluid liquid phases also a strongly correlated solid phase. Indirect excitons in semiconductor quantum wells have been found a favorable candidate due to their long-range pair interaction and the possibility to achieve strong nonideality by controlling the dipole moment with an external electric field. ased on first principle PIMC simulations we have computed the complete phase diagram in the region of the exciton crystal. (Quasi-)Long range crystalline order and macroscopic quantum coherence are found to be incompatible in an exciton crystal there is no supersolid phase, as long as the crystal is free of defects. The most striking feature of the crystal of indirect excitons is that it exhibits two quantum melting transitions which 10

persist at zero temperature: at low densities it melts by expansion whereas at high densities it melts when being compressed. The origin of this unusual and rich phase diagram has been traced to the non-trivial form of the exciton interaction potential. With it the exciton solid combines features of conventional neutral matter (exhibiting crystallization by compression) and Coulomb matter (quantum melting by compression), as found for instance in exotic compact stars. METHODS To verify the approximation (1) and the accuracy of the potential V xx we consider a two-exciton problem for fixed centers of mass, (R 0 1, R 0 2). Following Ref. [24] the exciton-exciton interaction energy is computed as the difference of the total energies of an exciton pair and of two single excitons, E XX (r hh ) = E(2X) 2E(X), where r hh = R 0 1 R 0 2. We solve the Schrödinger equation using the orn-oppenheimer approximation, neglecting hole exchange and, in a first approximation, assume infinitely heavy holes, m h. Then E XX = E e +1/r hh, where E e is the total energy of the two electrons which depends on their spin configuration. A first observation is that, in the symmetric state (singlet), E e, is not sensitive to r hh, once r hh d, see Fig. 1a. This is understood from the behavior of the electron density as a function of r hh (see Fig. 1b): in all cases the electron cloud extends well beyond r hh which is a result of the shallow interaction potential of an electron with the two holes, V eh = [(r + r hh /2) 2 + d 2 ] 1/2 [(r r hh /2) 2 + d 2 ] 1/2, for r hh < d, and the strong e-e repulsion evident from the pair distribution function g(r ee ), see Fig. 1c. Consequently, there is no noticeable difference of E e in the symmetric and antisymmetric states (see Fig.1a). With these results we can now analyze E XX (r hh ), cf. red dashed line in Fig. 1a. At large distances, r d, E XX practically coincides with the classical dipole potential, V D = d 2 /r 3, so we expect the system to behave like 2D polarized dipoles, at low densities. At smaller distances, r < d, however, E XX essentially follows a Coulomb potential which arises mainly from the holehole repulsion. Finally, for r d, the energy shows an unphysical Coulomb singularity which is due to the assumption of an infinite hole mass. Thus, we improved the model of Ref. [24] by taking into account a finite hole mass. This is found to have little influence on the total energy of the electrons, see red dots in Fig. 1a, as can be understood from the electron density and pair distribution, Fig. 1b,c. However, a finite hole mass strongly affects the interaction energy of two holes V hh and, consequently, the exciton interaction energy, which is now E XX = E e + V hh, red line in Fig. 1a. These results have been confirmed by independent multiconfiguration Hartree- 11

Fock simulations. Finally, we compute the exciton interaction energy using our PIMC simulations within the model 1 with the effective exciton potential V XX. The results for E XX are shown in Fig. 1a by the green squares and agree very well with solutions of the Schrödinger equation, confirming the quality of our model [Deviations are noticeable for r 5a which correspondso to densities which are outside the range of the present work]. Electronic address: bonitz@physik.uni-kiel.de [1] Anderson, M. H., Ensher, J. R., Matthews, M. R., Wieman, C. E. & Cornell, E. A. Observation of ose-einstein condensation in a dilute atomic vapor. Science 269, 198 201 (1995). [2] utov, L. V., Lai, C. W., Ivanov, A. L., Gossard, A. C. & Chemla, D. S. Towards ose-einstein condensation of excitons in potential traps. Nature 417, 47 52 (2002). [3] MacDonald, A. H. & Rezayi, E. H. Fractional quantum hall effect in a two-dimensional electron-hole fluid. Phys. Rev. 42, 3224 (1990). [4] Tiemann, L., Dietsche, W., Hauser, M. & von Klitzing, K. Critical tunneling currents in the regime of bilayer excitons. New J. Phys. 10, 045018 (2008). [5] Kasprzak, J. et al. ose-einstein condensation of exciton polaritons. Nature 443, 409 414 (2006). [6] Amo, A. et al. Superfluidity of polaritons in semiconductor microcavities. Nature Phys. 5, 805 810 (2009). [7] Lozovik, u.e. and erman, O.L. Phase transitions in a system of spatially separated electrons and holes. JETP 84, 1027 (1997) [8] Filinov, A. V., onitz, M. & Lozovik, Y. E. Excitonic clusters in coupled quantum dots. J. Phys. A 36, 5899 5904 (2003). [9] Filinov, A., onitz, M., Ludwig, P. & Lozovik, Y. E. Path integral Monte Carlo results for ose condensation of mesoscopic indirect excitons. phys. stat. sol. (c) 3, 2457 2460 (2006). [10] Timofeev, V.. & Gorbunov, A. V. Long-range coherence of interacting ose gas of dipolar excitons. J. Phys.: Condens. Matter 19, 295209 (2007). [11] Ludwig, P., Filinov, A. V., onitz, M. & Stolz, H. Quantum stark confined strongly correlated indirect excitons in quantum wells. phys. stat. sol. (b) 243, 2363 2366 (2006). [12] Domokos, P. & Ritsch, H. Collective cooling and Self-Organization of atoms in a cavity. Phys. Rev. Lett. 89, 253003 (2002). 12

[13] Gopalakrishnan, S., Lev,. L. & Goldbart, P. M. Emergent crystallinity and frustration with ose- Einstein condensates in multimode cavities. Nature Phys. 5, 845 850 (2009). [14] Griesmaier, A., Werner, J., Hensler, S., Stuhler, J. & Pfau, T. ose-einstein condensation of chromium. Phys. Rev. Lett. 94, 160401 (2005). [15] Sperlich, K. et al. Electric field-induced exciton localization in quantum wells. phys. stat. sol. (c) 6, 551 555 (2009). [16] Palo, S. D., Rapisarda, F. & Senatore, G. Excitonic condensation in a symmetric Electron-Hole bilayer. Phys. Rev. Lett. 88, 206401 (2002). [17] Astrakharchik, G. E., oronat, J., Kurbakov, I. L. & Lozovik, Y. E. Quantum phase transition in a Two-Dimensional system of dipoles. Phys. Rev. Lett. 98, 060405 (2007). [18] Filinov, A., Prokof ev, N. V. & onitz, M. erezinskii-kosterlitz-thouless transition in Two- Dimensional dipole systems. Phys. Rev. Lett. 105, 070401 (2010). [19] An alternative realization are two coupled n and p doped semiconductor layers. [20] At densities exceeding the Mott density n M pressure ionization transforms the system into an electronhole plasma. Here additional phases such as a hole liquid or a hole crystal are possible [31]. However, the present analysis is restricted to densities substantially below n M [21]. [21] S. en-tabou de-leon and. Laikhtman, Mott transition, biexciton crossover, and spin ordering in the exciton gas in quantum wells, Phys. Rev. 67, 235315 (2003). [22] Filinov, A., Ludwig, P., onitz, M. & Lozovik, Y. E. Effective interaction potential and superfluid solid transition of spatially indirect excitons. J. Phys. A 42, 214016 (2009). [23] Spatial separation of electrons and holes gives rise to spin polarization at low temperatures as was shown in Ref. [21]. [24] Schindler, C. and Zimmermann, R., Analysis of the exciton-exciton interaction in semiconductor quantum wells, Phys. Rev. 78, 045313 (2008) [25] Strandburg, K. J. Two-dimensional melting. Rev. Mod. Phys. 60, 161 (1988). [26] We also investigated the number of defects (excitons which do not have 6 nearest neighbors) and observed an abrupt increase at the melting temperature which is inconsistent with KTNHY theory, see supplementary figure S2. [27] Nelson, D. R. & Kosterlitz, J. M. Universal jump in the superfluid density of Two-Dimensional superfluids. Phys. Rev. Lett. 39, 1201 (1977). [28] Ceperley, D. M. Path integrals in the theory of condensed helium. Rev. Mod. Phys. 67, 279 (1995). 13

[29] üchler, H. P. et al. Strongly correlated 2D quantum phases with cold polar molecules: Controlling the shape of the interaction potential. Phys. Rev. Lett. 98, 060404 (2007). [30] Kalia, R. K. & Vashishta, P. Interfacial colloidal crystals and melting transition. J. Phys. C 14, L643 (1981). [31] onitz, M., Filinov, V. S., Fortov, V. E., Levashov, P. R. & Fehske, H. Crystallization in Two- Component Coulomb systems. Phys. Rev. Lett. 95, 235006 (2005). Acknowledgements We thank D. Hochstuhl for performing Multiconfiguration Hartree-Fock calculations for the exciton interaction energy. Stimulating discussions with Yu. Lozovik and P. Ludwig and financial support by the Deutsche Forschungsgemeinschaft (project FI 1252 and SF-TR24 project A5) are gratefully acknowledged. 14

Supplementary Material to the article Crystallization of an exciton superfluid J. öning, A. Filinov, and M. onitz Here we present details of our computations of the phase diagram of indirect excitons in two supplementary figures. Figure 6 illustrates the computation of the winding number versus temperature (left) and the finite size scaling for the critical temperature of the KT transition (right). The scaling allows us to extrapolate to an infinite system, L and thus make predictions for a macroscopic system. 15

W 2 (T ) 1.7 1.6 1.5 1.4 1.3 (a) 170 na 2 N=56 = 5 10 3 [ktc/ha, 10 3 ] 1.2 1.45 1.5 1.55 1.6 1.65 1.7 1.75 [k T /Ha, 10 3 ] 2.2 2 1.8 1.6 1.4 (b) T c /4 T c /2 na 2 10 2 5 10 3 = 2 10 2 0 0.05 0.1 1/ ln 2 L Figure 6: (a) Temperature dependence of the winding number W 2 (T ) for two exciton numbers N = 56 and 170 at a density of na 2 = 5 10 3. The erezinskii-kosterlitz-thouless transition temperature T c (N) is determined by the condition [1], W 2 (T c (N)) = 4/π, shown by the horizontal dashed line. observes a systematic shift of T c (N) to lower values with the increase of the system size N. (b) System size dependence of T c (L = N/n) for three densities: na 2 One = 5 10 3, 10 2, 2 10 2. The extrapolation to the thermodynamic limit, T c (L ), is obtained by fitting the simulation data by the equation: T c (L) = T c ( ) + b/ ln 2 (L). It is a direct consequence of the Kosterlitz-Thouless renormalization group analysis [1] which is considered to be exact in the asymptotic regime of large L. The second figure explores the nature of the melting transition at constant density. We analyze the number of defects, i.e. the number of particles (the probability) not having six nearest neighbors. There is a sharp increase of the number of defects at the melting point which is in disagreement with the KTNHY scenario. A possible alternative to the KTNHY thery of melting in 2D is a first order solid-liquid phase transition, with an exponential decay of g 6 (r). This is not observed in our simulations, possibly due to a limited system size. In the simulations, we observe 16

an accumulation of the defects at grain boundaries formed between small crystallites. This, as was recently verified for classical 2D systems, forces the phase transition to be of first order [2]. To answer this question for the present system would require extensive additional simulations. 0.45 0.425 0.4 1 P6 0.375 0.35 0.325 0.3 0.6 0.8 1 1.2 1.4 1.6 1.8 2 Temperature, [k T /Ha, 10 3 ] Figure 7: Temperature dependence of the defect fraction at density na 2 = 2.2 10 3 and particle numbers N = 501 505 (vertically aligned dots), which characterizes the ordered and disordered phases. A Voronoi analysis was performed to provide access to local distortions of the hexagonal symmetry which are characterized by 1 P 6 where P 6 is the probability that a particle has 6 nearest neighbors. The defect fraction shows a sharp jump at T 10 3 Ha. This is in disagreement with the KTNHY theory, which predicts a continuous unbinding of dislocations in the hexatic phase indicated by a continuous variation of the critical exponent η 6 (T ) 1/4 and the bond angular correlation function g 6 (r) r η 6(T ). In contrast, we observe an abrupt transition to quasi-long-range angular order with η 6 (T ) 2, cf. Fig. 2. of the main article. Electronic address: bonitz@physik.uni-kiel.de [1] D. R. Nelson and J. M.Kosterlitz, Phys. Rev. Lett. 39, 1201 (1977) [2] P. Hartmann et. al., Phys. Rev. Lett. 105, 115004 (2010) 17