<<

week ending PRL 106, 030401 (2011) PHYSICAL REVIEW LETTERS 21 JANUARY 2011

Thermodynamics of the 3D Hubbard Model on Approaching the Ne´el Transition

Sebastian Fuchs,1 Emanuel Gull,2 Lode Pollet,3 Evgeni Burovski,4,5 Evgeny Kozik,3 Thomas Pruschke,1 and Matthias Troyer3 1Institut fu¨r Theoretische Physik, Georg-August-Universita¨tGo¨ttingen, 37077 Go¨ttingen, Germany 2Department of Physics, Columbia University, New York, New York 10027, USA 3Theoretische Physik, ETH Zurich, 8093 Zurich, Switzerland 4LPTMS, CNRS and Universite´ Paris-Sud, UMR8626, Baˆtiment 100, 91405 Orsay, France 5Department of Physics, Lancaster University, Lancaster, LA1 4YB, United Kingdom (Received 24 September 2010; published 18 January 2011) We study the thermodynamic properties of the 3D Hubbard model for temperatures down to the Ne´el temperature by using cluster dynamical mean-field theory. In particular, we calculate the energy, entropy, density, double occupancy, and nearest-neighbor correlations as a function of chemical potential, temperature, and repulsion strength. To make contact with cold-gas experiments, we also compute properties of the system subject to an external trap in the local density approximation. We find that an entropy per particle S=N 0:65ð6Þ at U=t ¼ 8 is sufficient to achieve a Ne´el state in the center of the trap, substantially higher than the entropy required in a homogeneous system. Precursors to antiferromagnetism can clearly be observed in nearest-neighbor spin correlators.

DOI: 10.1103/PhysRevLett.106.030401 PACS numbers: 05.30.Fk, 03.75.Ss, 71.10.Fd

The Hubbard model remains one of the cornerstone the entropy, energy, density, double occupancy, and spin models in , capturing the essence correlations—for interactions U up to the bandwidth 12t of strongly correlated electron physics relevant to high- on approach to the Ne´el temperature TN by performing temperature superconductors [1] and correlation-driven controlled large-scale cluster dynamical mean-field calcu- insulators [2]. While qualitative features of the phase lations and extrapolations to the infinite system size limit, diagram are known from analytical approximations, con- as well as determinantal diagrammatic Monte Carlo trolled quantitative studies in the low-temperature regimes (DMC) simulations at half filling. We use this information relevant for applications are not readily tractable with tools to calculate the entropy per particle required for experi- presently available. A recent program that aims to imple- ments on ultracold atomic gases in optical lattices to reach ment the Hubbard model in a cold gases experiment [3] has aNe´el state in the trap center. We finally show that the led to experimental signs of the Mott [4,5]. nearest-neighbor spin-correlation function contains clear Modeling by the dynamical mean-field theory (DMFT) precursors for antiferromagnetism that may already be [5,6] and high-temperature series expansions (HTSE) [7] detectable in current generation experiments and that resulted in temperature and entropy estimates [8]. A major are useful for thermometry (more so than measurements experimental achievement will be the detection of the of the double occupancy) close to TN. antiferromagnetic phase, for which the slow and ill- The Hubbard model is defined by its Hamiltonian understood equilibration rates, the limited number of de- X X X y tection methods, and inherent cooling problems will have H^ ¼t c^ic^j þ U n^i"n^i# in^i; (1) to be overcome. hi;ji; i i; Experimental progress has also sparked interest in simu- cy ¼"; # lations of the 3D Hubbard model, where new algorithms, where ^i creates a with spin component y such as the real-space DMFT [9,10] and diagrammatic on site i, n^i ¼ c^ic^i, h...i denotes summation over Monte Carlo [11] methods, have been developed. As in neighboring lattice sites, t is the hopping amplitude, the case of , where synergy between experiment and U is the on-site repulsion, and i ¼ Vðr~iÞ with simulation has led to quantitative understanding of experi- the chemical potential and Vðr~iÞ the confining potential ments [12], accurate results for the of at the location of the ith lattice site. We set Vðr~Þ¼0 in all the 3D Hubbard model will be important for validation, calculations and consider realistic traps later on. calibration, and thermometry of fermionic experiments. A Our numerical approach is a cluster generalization of the crucial role is played by the entropy, since these experi- DMFT [13]. In the cluster DMFT the self-energy is ap- N ments form isolated systems where parameters are changed proximated by cPmomentum-dependent basis functions Nc adiabatically, not isothermally. KðkÞ: ðk; !Þ K KðkÞKð!Þ. The exact problem is In this Letter, we provide the full thermodynamical recovered for Nc !1. Within the dynamical cluster ap- equation of state of the Hubbard model—in particular, proximation (DCA) [14] used here, KðkÞ are piecewise

0031-9007=11=106(3)=030401(4) 030401-1 Ó 2011 American Physical Society week ending PRL 106, 030401 (2011) PHYSICAL REVIEW LETTERS 21 JANUARY 2011 constant over momentum patches, and DMFT [15] corre- sponds to Nc ¼ 1 and ¼ ð!Þ. Solving the DMFT and DCA equations requires the solution of a quantum impurity model. Continuous- time quantum impurity solvers [16–18], in particular, the continuous-time auxiliary field method [17] with submatrix updates [19] used here, have made it possible to solve such models efficiently and numerically exactly on large clus- ters, thereby providing a good starting point for an extrapo- lation of finite-size clusters to the infinite system [11,20]We have performed extensive DCA calculations on bipartite clusters with Nc ¼ 18, 26, 36, 48, 56, and 64. In order to achieve an optimal scaling behavior we exclusively use the clusters determined in Ref. [20] following the criteria pro- posed by Ref. [21]. As the DCA exhibits a 1=L2 finite-size 1=3 scaling in the linear cluster size L ¼ Nc away from criti- cal behavior [22], we extrapolate our cluster results linearly s 2=3 FIG. 1 (color online). Entropy per lattice site of the Hubbard in Nc . DCA error bars include extrapolation uncertain- model as a function of temperature T=t, for U=t ¼ 8, at half filling. ties; a sample extrapolation is given in Ref. [23]. Despite a Dashed vertical lines (black): TN from Ref. [20]; dotted lines sign problem away from half filling, temperatures T=t (blue): according to DMC simulations. Dashed horizontal lines 0:4, on the order of the Ne´el temperature, are reliably (black): entropy per lattice site s at TN [20]; dotted lines (blue): accessible for all but the largest interaction strength according to DMC simulations; logð2Þ is shown as a solid (green) U ¼ 12t where we have been restricted to T=t 0:5. horizontal line. Inset: Energy E=t per lattice site. The DMC data The potential energy, double occupancy, and nearest- were extrapolated linearly in 1=L from the data at L ¼ 6; 8; 10. neighbor spin-spin correlation haveP been measured E ¼ ðk~ÞGðk;~ i! Þ The double occupancy, which has played a crucial role directly. The kinetic energy kin n;k~ n has ~ in optical lattice experiments [4,5,7,26], is shown in Fig. 2 been calculated by summing ðkÞ, the bare dispersion as a function of temperature at half filling for different of the simple cubic lattice, and the single-particle Green values of U=t. While for small U=t a remarkable increase ~ ~ function Gðk; i!nÞ over all momenta k and Matsubara is seen on approach to TN, only a plateau remains at frequencies i!n. The entropy S has subsequently been moderate values of U=t. This is in contrast to the DMFT calculated by numeric integration: predictions but similar to lattice quantum Monte Carlo EðT Þ EðTÞ Z T EðT0Þ results in two dimensions [27]. For larger interactions SðTÞ¼SðT Þ u þ u dT0 ; u 0 (2) (U=t * 12), the double occupancy rises above that of a T T T T 2 u single-site paramagnet, consistent with DMFT results for up to a Tu=t 10 , where the entropy SðTuÞ is accurately given by HTSE. Tables of the complete results containing finite cluster and extrapolated values at and away from half filling for the entropy, energy, density, double occupancy, and spin correlations are given in the supplementary ma- terial [23]. Results at half filling.—We start our analysis at half filling and focus on U=t ¼ 8, where a comparison with results from lattice simulations [24,25] is possible. We see in Fig. 1 that the entropy calculated by using the DCA and DMC simulations coincides within error bars at all tem- peratures. Agreement with a 10th order high-temperature series expansion [7] is found down to T=t 1:6. At that temperature also the single-site DMFT starts to deviate because that method misses short-range antiferromagnetic correlations. The Ne´el temperature was found to be TN=t 0:36ð2Þ in Ref. [20]. Our DMC calculations find it at TN=t ¼ 0:333ð7Þ. By using the DCA, the critical s : ð Þ T entropy is 0 46 4 for N according to Ref. [20], and FIG. 2 (color online). Double occupancy of the Hubbard s :¼ S=Nc 0:42ð2Þ with TN according to the DMC model as a function of T=t, at half filling. Extrapolated DCA simulations. We will use DMC numbers for TN in the results are shown as solid lines and DMFT values as dashed rest of this Letter. lines. Vertical lines: See Fig. 1. 030401-2 week ending PRL 106, 030401 (2011) PHYSICAL REVIEW LETTERS 21 JANUARY 2011

that the entropy per particle number N increases strongly at lower densities. While the single-site DMFT remains accurate for densities n & 0:6 due to the weak momentum dependence of the self-energy in this regime [30], the DCA results are important closer to half filling. Similarly, near half filling the DMFT overestimates the double occupancy by 10%, while deviations are less pronounced at lower densities. This observation persists for all interactions and temperatures investigated. The flattening of the double occupancy for 8 U=t 12 (cf. Fig. 2) is also seen away from half filling and leads to virtually unchanged profiles over the trap in an optical lattice system. On the other hand, the spin-spin-correlation function away from half filling (inset in Fig. 3) changes most rapidly near half filling when approaching TN since it couples strongly to the developing (short-range) spin correlations. Entropy in the optical lattice system.—We now turn to FIG. 3 (color online). Nearest-neighbor spin-spin correlation the experimentally relevant case of an optical lattice in a of the Hubbard model as a function of T=t, at half filling. Inset: Density dependence for U=t ¼ 8 and selected temperatures. harmonic trap, which is a closed system where entropy is Vertical lines: See Fig. 1. conserved. We choose parameters close to current experi- ments: Vðr~Þ¼0:004ðjr~j=aÞ2t with lattice spacing a, and we consider the case of half filling in the trap center: the antiferromagnetic phase below TN [10]. The negative ¼ U= slope of DðTÞ, discussed in the context of the single-site 2. We treat the harmonic confinement in a local DMFT [28], persists for a wide range of parameters. density approximation (LDA): For every site we perform a DCA simulation for a homogeneous system and average Sharp features just above TN, as detected in single-site (momentum-independent) studies [10], are not observed the results over the trap. LDA was found to be a good for the interaction values and temperature ranges studied approximation for the Bose-Hubbard model for wide traps, here. Hence the proposal that the double occupancy is a except in close proximity to the critical point [31–33]of Uð Þ good candidate for thermometry is not substantiated by the 1 phase transition because of the diverging corre- more accurate momentum-dependent calculations. lation length. In our setup LDA errors are small compared T The spin-spin-correlation function, plotted in Fig. 3 as a to errors from the uncertainty of N. function of temperature for various U and as a function of Because of the large volume fraction, the wings of the filling for T=t ¼ 0:3, 0.8, and 1.6 at U=t ¼ 8, is accessible gas may capture more entropy than the center of the trap, only in methods that include nonlocal correlations but even though the entropy per site is comparable to the one in the center (see Fig. 5). In fact, the entropy of the may be accessible experimentally [29]. It has a steep slope :

FIG. 5 (color online). Entropy profiles (entropy per lattice site) FIG. 4 (color online). Entropy per lattice site s and the entropy plotted over the trap in the LDA approximation for different per particle S=N (inset) of the Hubbard model at the temperature temperatures with an interaction strength U=t ¼ 8. Error bars, T ¼ 0:35t TN, as a function of density n, for U=t ¼ 8. shown every 5 lattice spacings, are smaller than symbol size. 030401-3 week ending PRL 106, 030401 (2011) PHYSICAL REVIEW LETTERS 21 JANUARY 2011

and L. Tarruell. This work was supported by the Swiss National Science foundation, the National Science Foundation Grants No. DMR-0705847 and No. PHY- 0653183, ANR Grant No. ANR-BLAN-6238, the Aspen Center for Physics, a grant from the Army Research Office with funding from the DARPA OLE program, and by the DFG through the collaborative research center SFB 602. We used the Brutus cluster at ETH Zurich and Norddeutscher Verbund fu¨r Hoch- und Ho¨chstleistungsrechnen. Simulation codes were based on Algorithms and Libraries for Physics Simulations [34]. While finalizing this Letter, we became aware of a related effort [35] using the DMFT and HTSE.

FIG. 6 (color online). Entropy per particle averaged over the trap as a function of temperature relative to TN for different U. The Ne´el temperature is reached at T=t ¼ 0:333ð7Þ for U=t ¼ 8, [1] P. W. Anderson, Science 235, 1196 (1987). when the average entropy is S=N ¼ 0:65ð6Þ. Errors (not shown) [2] M. Imada, A. Fujimori, and Y. Tokura, Rev. Mod. Phys. in SðTNÞ are estimated to be in the 10% range, with the largest 70, 1039 (1998). contribution caused by the uncertainty in TN. [3] M. Ko¨hl et al., Phys. Rev. Lett. 94, 080403 (2005). [4] R. Jo¨rdens et al., Nature (London) 455, 204 (2008). [5] U. Schneider et al., Science 322, 1520 (2008). trap which is about 50% larger than what could be expected [6] L. De Leo et al., Phys. Rev. Lett. 101, 210403 (2008). from a homogeneous study. Optimal parameters are [7] V.W. Scarola et al., Phys. Rev. Lett. 102, 135302 (2009). U=t ¼ T =t ¼ : ð Þ around 8 when N 0 333 7 according to [8] R. Jo¨rdens et al., Phys. Rev. Lett. 104, 180401 (2010). DMC simulations, corresponding to S=N ¼ 0:65ð6Þ in [9] M. Potthoff and W. Nolting, Phys. Rev. B 59, 2549 (1999). the trap, while S=N ¼ 0:42ð2Þ would be expected for a [10] E. V. Gorelik et al., Phys. Rev. Lett. 105, 065301 (2010). homogeneous system. As seen in Fig. 6, all U in the range [11] E. Kozik et al., Europhys. Lett. 90, 10 004 (2010). 8 < U=t < 12 lead to similar conclusions. We have [12] S. Trotzky et al., Nature Phys. 6, 998 (2010). verified that changing the trap by a factor of 4 does not [13] T. Maier et al., Rev. Mod. Phys. 77, 1027 (2005). alter these conclusions. [14] M. H. Hettler et al., Phys. Rev. B 58, R7475 (1998). Conclusions.—We have provided the full thermody- [15] A. Georges et al., Rev. Mod. Phys. 68, 13 (1996). [16] A. N. Rubtsov, V.V. Savkin, and A. I. Lichtenstein, Phys. namics of the 3D Hubbard model in the thermodynamic U=t Rev. B 72, 035122 (2005). limit by using the DCA formalism for 12 and [17] E. Gull et al., Europhys. Lett. 82, 57 003 (2008). temperatures above the Ne´el temperature. Comparing to [18] P. Werner et al., Phys. Rev. Lett. 97, 076405 (2006). single-site DMFT results we found that the latter already [19] E. Gull et al., arXiv:1010.3690. fail at remarkably high temperatures (T=t 1:5 for [20] P. Kent et al., Phys. Rev. B 72, 060411(R) (2005). U=t ¼ 8 at the 1% level) near half filling. While the [21] D. D. Betts and G. E. Stewart, Can. J. Phys. 75, 47 (1997). entropy per particle at the Ne´el temperature TN=t ¼ [22] T. A. Maier and M. Jarrell, Phys. Rev. B 65, 041104 (2002). 0:333ð7Þ (determined with DMC simulations) is [23] See supplemental material at http://link.aps.org/ S=N ¼ 0:42ð2Þ for U=t ¼ 8 in a homogeneously half filled supplemental/10.1103/PhysRevLett.106.030401 for all system, we find that the Ne´el transition in a trap can raw data of the 3D Hubbard model computed with various methods for several interaction strengths. already be reached at S=N ¼ 0:65ð6Þ in a realistically T [24] R. Staudt, M. Dzierzawa, and A. Muramatsu, Eur. Phys. J. sized harmonic trap (taking N according to Ref. [20] leads B 17, 411 (2000). S=N ¼ : to 0 69). [25] E. Burovski et al., Phys. Rev. Lett. 101, 090402 (2008). We have also investigated the double occupancy and [26] N. Strohmaier et al., Phys. Rev. Lett. 104, 080401 (2010). the nearest-neighbor spin-spin-correlation function as ex- [27] T. Paiva et al., Phys. Rev. Lett. 104, 066406 (2010). perimentally measurable quantities that were suggested to [28] F. Werner et al., Phys. Rev. Lett. 95, 056401 (2005). show precursors of antiferromagnetism. The double occu- [29] S. Trotzky et al., Phys. Rev. Lett. 105, 265303 (2010). pancy is almost flat as a function of temperature, while the [30] E. Gull et al., Phys. Rev. B 82, 155101 (2010). spin correlations show a strong temperature dependence [31] S. Wessel et al., Phys. Rev. A 70, 053615 (2004). around the Ne´el temperature. This suggests that the spin [32] M. Campostrini and E. Vicari, Phys. Rev. Lett. 102, 240601 (2009); Phys. Rev. A 81, 023606 (2010). correlations, not the double occupancy, are best suited to [33] L. Pollet, N. V. Prokof’ev, and B. V. Svistunov, Phys. Rev. observe precursors of antiferromagnetism and measure the Lett. 104, 245705 (2010). temperature. Our numerical data can be used to calibrate [34] A. F. Albuquerque et al., J. Magn. Magn. Mater. 310, 1187 such a spin-correlation thermometer. (2007). We acknowledge stimulating discussions with I. Bloch, [35] L. De Leo et al., arXiv:1009.2761 [Phys. Rev. A (to be T. Esslinger, A. Georges, D. Greif, O. Parcollet, V. Scarola, published)]. 030401-4