thermal conductivity of the one-dimensional fermi-hubbard model

11
Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model C. Karrasch, 1,2,5 D. M. Kennes, 3 and F. Heidrich-Meisner 4 1 Department of Physics, University of California, Berkeley, California 95720, USA 2 Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA 3 Institut für Theorie der Statistischen Physik, RWTH Aachen University and JARA-Fundamentals of Future Information Technology, 52056 Aachen, Germany 4 Department of Physics and Arnold Sommerfeld Center for Theoretical Physics, Ludwig-Maximilians-Universität München, 80333 München, Germany 5 Dahlem Center for Complex Quantum Systems and Fachbereich Physik, Freie Universität Berlin, 14195 Berlin, Germany (Received 18 June 2015; published 9 September 2016) We study the thermal conductivity of the one-dimensional Fermi-Hubbard model at a finite temperature using a density matrix renormalization group approach. The integrability of this model gives rise to ballistic thermal transport. We calculate the temperature dependence of the thermal Drude weight at half filling for various interaction strengths. The finite-frequency contributions originating from the fact that the energy current is not a conserved quantity are investigated as well. We report evidence that breaking the integrability through a nearest-neighbor interaction leads to vanishing Drude weights and diffusive energy transport. Moreover, we demonstrate that energy spreads ballistically in local quenches with initially inhomogeneous energy density profiles in the integrable case. We discuss the relevance of our results for thermalization in ultracold quantum-gas experiments and for transport measurements with quasi-one- dimensional materials. DOI: 10.1103/PhysRevLett.117.116401 Improving our understanding of transport in one- dimensional (1D) strongly correlated systems (SCSs) is an active field in condensed matter theory. While in 1D the existence of powerful numerical [13] and analytical [4,5] methods makes it possible to obtain quantitative results (see, e.g., [4,68]), transport coefficients are very hard to come by exactly and are challenging quantities to deter- mine numerically. Early studies suggested that integrable systems such as the 1D spin-1=2 Heisenberg or Fermi- Hubbard model (FHM) may possess ballistic transport properties at finite temperatures [9]. In the linear response theory, ballistic dynamics manifests itself through nonzero Drude weights. The so far best understood model is the spin-1=2 XXZ chain, for which the Drude weight for thermal transport has been calculated exactly [10,11], while substantial progress has recently been made regarding the spin conductivity [1221]. The theory of transport in the 1D FHM is much less advanced and has focused on spin and charge transport [2228]. The thermal conductivity can, by using the Kubo formula [29,30], be written as Re κðωÞ¼ 2πD th ðT ÞδðωÞþ κ reg ðωÞ ð1Þ with the thermal Drude weight D th ðT Þ and a regular finite-frequency contribution κ reg ðωÞ. Since in SCSs the Wiedemann-Franz law is not necessarily valid, indepen- dent calculations of charge and thermal transport are required. The formal argument to prove a nonzero Drude weight relies on the Mazur inequality [9,31,32] D th 1 2T 2 L X i hQ i I th i 2 hQ 2 i i ; ð2Þ where I th is the energy-current operator and the Q i are local or quasilocal conserved quantities [12,15,33]. In the presence of interactions, nontrivial Q i leading to finite Drude weights typically exist in integrable models [9]. For instance, for the spin-1=2XXZ chain, the energy current I th ¼ Q 3 itself is conserved [implying that κ reg ðωÞ¼ 0], while for the FHM, I th has only a partial overlap with Q 3 (both operators have a similar structure [34]) such that while D th > 0, also κ reg ðωÞ 0 [9]. As a consequence, the half-filled FHM realizes an unusual behavior: ballistic thermal transport [9], yet diffusive charge conduction [27,28] at temperatures T> 0. Here, we address the outstanding problem of quantita- tively calculating Re κðωÞ at T> 0 for the FHM at half filling by using a finite-temperature density matrix renorm- alization group (DMRG) method [3,4751], previously applied to both charge transport in the FHM [27] and transport in quasi-1D spin-1=2 systems [14,21,47,52,53]. We obtain the energy-current autocorrelation function C th ðtÞ¼ RehI th ðtÞI th i=L from time-dependent simulations. Since C th ðtÞ saturates fast at a time-independent nonzero value, we are able to extract (i) the thermal Drude weight and (ii) the regular part from a Fourier transformation in combination with a linear prediction [54]. Moreover, PRL 117, 116401 (2016) PHYSICAL REVIEW LETTERS week ending 9 SEPTEMBER 2016 0031-9007=16=117(11)=116401(7) 116401-1 © 2016 American Physical Society

Upload: trinhnguyet

Post on 14-Feb-2017

224 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

C. Karrasch,1,2,5 D. M. Kennes,3 and F. Heidrich-Meisner41Department of Physics, University of California, Berkeley, California 95720, USA

2Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA3Institut für Theorie der Statistischen Physik, RWTH Aachen University and JARA-Fundamentals of Future Information Technology,

52056 Aachen, Germany4Department of Physics and Arnold Sommerfeld Center for Theoretical Physics, Ludwig-Maximilians-Universität München,

80333 München, Germany5Dahlem Center for Complex Quantum Systems and Fachbereich Physik, Freie Universität Berlin, 14195 Berlin, Germany

(Received 18 June 2015; published 9 September 2016)

We study the thermal conductivity of the one-dimensional Fermi-Hubbard model at a finite temperatureusing a density matrix renormalization group approach. The integrability of this model gives rise to ballisticthermal transport. We calculate the temperature dependence of the thermal Drude weight at half filling forvarious interaction strengths. The finite-frequency contributions originating from the fact that the energycurrent is not a conserved quantity are investigated as well. We report evidence that breaking theintegrability through a nearest-neighbor interaction leads to vanishing Drude weights and diffusive energytransport. Moreover, we demonstrate that energy spreads ballistically in local quenches with initiallyinhomogeneous energy density profiles in the integrable case. We discuss the relevance of our results forthermalization in ultracold quantum-gas experiments and for transport measurements with quasi-one-dimensional materials.

DOI: 10.1103/PhysRevLett.117.116401

Improving our understanding of transport in one-dimensional (1D) strongly correlated systems (SCSs) isan active field in condensed matter theory. While in 1D theexistence of powerful numerical [1–3] and analytical [4,5]methods makes it possible to obtain quantitative results(see, e.g., [4,6–8]), transport coefficients are very hard tocome by exactly and are challenging quantities to deter-mine numerically. Early studies suggested that integrablesystems such as the 1D spin-1=2 Heisenberg or Fermi-Hubbard model (FHM) may possess ballistic transportproperties at finite temperatures [9]. In the linear responsetheory, ballistic dynamics manifests itself through nonzeroDrude weights. The so far best understood model is thespin-1=2 XXZ chain, for which the Drude weight forthermal transport has been calculated exactly [10,11], whilesubstantial progress has recently been made regarding thespin conductivity [12–21].The theory of transport in the 1D FHM is much less

advanced and has focused on spin and charge transport[22–28]. The thermal conductivity can, by using the Kuboformula [29,30], be written as

Re κðωÞ ¼ 2πDthðTÞδðωÞ þ κregðωÞ ð1Þ

with the thermal Drude weight DthðTÞ and a regularfinite-frequency contribution κregðωÞ. Since in SCSs theWiedemann-Franz law is not necessarily valid, indepen-dent calculations of charge and thermal transport arerequired.

The formal argument to prove a nonzero Drude weightrelies on the Mazur inequality [9,31,32]

Dth ≥1

2T2L

Xi

hQiIthi2hQ2

i i; ð2Þ

where Ith is the energy-current operator and the Qi arelocal or quasilocal conserved quantities [12,15,33]. In thepresence of interactions, nontrivial Qi leading to finiteDrude weights typically exist in integrable models [9]. Forinstance, for the spin-1=2XXZ chain, the energy currentIth ¼ Q3 itself is conserved [implying that κregðωÞ ¼ 0],while for the FHM, Ith has only a partial overlap with Q3

(both operators have a similar structure [34]) such thatwhile Dth > 0, also κregðωÞ ≠ 0 [9]. As a consequence, thehalf-filled FHM realizes an unusual behavior: ballisticthermal transport [9], yet diffusive charge conduction[27,28] at temperatures T > 0.Here, we address the outstanding problem of quantita-

tively calculating Re κðωÞ at T > 0 for the FHM at halffilling by using a finite-temperature density matrix renorm-alization group (DMRG) method [3,47–51], previouslyapplied to both charge transport in the FHM [27] andtransport in quasi-1D spin-1=2 systems [14,21,47,52,53].We obtain the energy-current autocorrelation functionCthðtÞ ¼ RehIthðtÞIthi=L from time-dependent simulations.Since CthðtÞ saturates fast at a time-independent nonzerovalue, we are able to extract (i) the thermal Drude weightand (ii) the regular part from a Fourier transformationin combination with a linear prediction [54]. Moreover,

PRL 117, 116401 (2016) P HY S I CA L R EV I EW LE T T ER Sweek ending

9 SEPTEMBER 2016

0031-9007=16=117(11)=116401(7) 116401-1 © 2016 American Physical Society

Page 2: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

we consider the extended FHM as an example for anonintegrable model and provide evidence that ballisticcontributions are absent, with a diffusive form of thelow-frequency κregðωÞ.The FHM has been realized with ultracold quantum

gases [55–64]. In ultracold quantum gases, relaxationprocesses play an important role for reaching thermalequilibrium during the state preparation [65,66], andthermometry is an open experimental problem [67].Furthermore, understanding thermalization dynamics andnonequilibrium transport as such have been the goal ofseveral optical-lattice experiments with Hubbard systems[68–71]. We demonstrate that real-space perturbations inthe energy density spread ballistically in the 1D FHMat T > 0 while charge diffuses [27,28], providing a routeto experimentally observing the qualitative differencebetween charge and energy dynamics in this model.Definitions.—The Hamiltonian of the extended FHM is

given by H ¼ PL−1l¼1 hl with local terms

hl ¼ −t0Xσ

ðc†lσclþ1σ þ H:c:Þ þ Vðnl − 1Þðnlþ1 − 1Þ

þU2

��nl↑ −

1

2

��nl↓ −

1

2

þ�nlþ1↑ −

1

2

��nlþ1↓ −

1

2

��; ð3Þ

where clσ annihilates a fermion with spin σ on site l andnlσ ¼ c†lσclσ. U and V denote the on-site and the nearest-neighbor Coulomb repulsion, respectively. We use openboundary conditions. All results in the main text are forhalf filling n ¼ N=L ¼ 1, where N is the total number offermions. For convenience, we implemented the FHM as atwo-leg spin-1=2 ladder [34].We derive the energy current from the continuity equation

[9], leading to Ith ¼ iP

L−2l¼1 ½hl; hlþ1� (for the full expression,

see [34]). At n ¼ 1, particle-hole symmetry leads to avanishing thermopower [34,72], and, thus, the thermalconductivity stems solely from energy-current correlations.Numerical method.—The thermal Drude weight is

related to the long-time asymptote of the current correlationfunctions via

Dth ¼ limt→∞

limL→∞

CthðtÞ2T2

; ð4Þ

and the regular part of the conductivity defined in Eq. (1)can be obtained from [ ~CthðtÞ ¼ CthðtÞ − 2T2Dth]

κregðωÞ ¼1 − e−ω=T

ωTRe

Z∞

0

dteiωt limL→∞

~CthðtÞ: ð5Þ

Note that the derivation of the Kubo formula for κ ismore subtle than for the charge conductivity (see, e.g.,Refs. [30,73–77], and references therein, and Ref. [34] for

a discussion), while there is also ongoing research onthermal and energy transport in an open quantum system(see, e.g., [78–81]).Our finite-T DMRG method, implemented via matrix-

product states [82–85], is based on the purification trick[86] (see [49,54,87–89] for related work). Thus, wesimulate pure states that live in a Hilbert space spannedby the physical and auxiliary (ancilla) degrees of freedom.Mixed states are obtained by tracing over the ancillas. Inorder to access time scales as large as possible, we employ afinite-temperature disentangler [47], using that purificationis not unique to slow down the entanglement growth.Moreover, we exploit “time-translation invariance” [49],rewrite hIthðtÞIthð0Þi ¼ hIthðt=2ÞIthð−t=2Þi, and carry outtwo independent calculations for Ithðt=2Þ as well as forIthð−t=2Þ. Since the energy current is a six-point function, atDMRG simulation for the energy-current autocorrelationis much more demanding than it is in the charge case. Ourcalculations are performed with L ¼ 100 sites (see Fig. S1in Ref. [34] for an analysis of the L dependence). The“finite-time” error of κregðωÞ can be assessed followingRef. [53], resulting in the error bars shown in the figures.Real-time decay of Cth.—Typical DMRG results for

CthðtÞ are shown in Figs. 1(a) and 1(b) for U ¼ 2t0 andU ¼ 8t0, respectively, and temperatures T ¼ ∞, 2t0, 2t0=3.We are able to reach times tt0 ≲ 5. For V ¼ 0 (thick lines),i.e., in the integrable case, CthðtÞ rapidly saturates at aconstant nonzero value, reflecting the ballistic nature ofenergy transport in this model. The transients are surpris-ingly short compared to spin transport in the spin-1=2 XXZchain [14] and exhibit oscillations with a fairly smallamplitude. To illustrate the behavior in the nonintegrableextended FHM, we present data for V ¼ U=3 (thin lines),for which our system Eq. (3) is still in the Mott-insulatingphase [90]. The DMRG results unveil a much stronger

Cth

(t)/

t 04

0 2 4t t0

0

6

12

0

1

2

T/t0=∞T/t0=2T/t0=0.66

(a) U/t0=2

(b) U/t0=8 thick: V=0thin : V=U/3

2T2*Dth

x10

V=0

V=0

V=0

V=0

FIG. 1. DMRG results for CthðtÞ for various T > 0 at(a) U=t0 ¼ 2 and (b) U=t0 ¼ 8. Thick lines, V ¼ 0; thin lines,V ¼ U=3. The curves for U=t0 ¼ 8 and T=t0 ¼ 0.66 are multi-plied by a factor of 10.

PRL 117, 116401 (2016) P HY S I CA L R EV I EW LE T T ER Sweek ending

9 SEPTEMBER 2016

116401-2

Page 3: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

decay of CthðtÞ compared to the integrable case yet alsolonger transient dynamics before the asymptotic regime isreached [see, e.g., the data for T ¼ ∞ shown in Fig. 1(a)].For U=t0 ¼ 8 and T ¼ ∞, the real-time decay of Cth isconsistent with a vanishing DthðTÞ, as expected for thisnonintegrable model [53,91–94].Thermal Drude weight for V ¼ 0.—The fast saturation

of CthðtÞ at a constant and nonzero value allows us toextract the temperature dependence of DthðTÞ, displayed inFig. 2 for U=t0 ¼ 0, 1, 2, 4, 8 (note the log-log scale). ForU ¼ 0, we compare our data to the exact result

DthðTÞ ¼t20

2πT2

−π½ϵkvkfðϵkÞ�2eϵk=Tdk; ð6Þ

where ϵk ¼ −2t0 cosðkÞ, vk ¼ ∂ϵk=∂k, and fðϵÞ ¼1=ð1þ eϵ=TÞ. The agreement is excellent. In general,DthðTÞ has a maximum at a U-dependent temperature thatshifts to a larger temperature as U increases. In the high-temperature regime T > t0, Dth ¼ D∞

th=T2 (dashed lines in

the figure), where the prefactorD∞th has been extracted from

the numerical data at β ¼ 0. In the Supplemental Material[34], we compare the thermal Drude weight of the FHM tothe one of the Heisenberg chain [10]. The latter describesthe low-temperature contribution of spin excitations to thefull Dth of the FHM for U ≫ t0 at n ¼ 1, while we presentresults for the spin-incoherent regime T ≫ 4t20=U, wherecharge excitations dominate.We next study howmuch of the full spectral weight of Re

κðωÞ is in the Drude peak by plotting 2πDth=I0 versusU=t0at β ¼ 0 in the inset in Fig. 2, where I0 ¼

RdωReκðωÞ. The

Drude weight contains the full weight I0 only at U ¼ 0 andfor U=t0 → ∞. In the former case, this results from the

exact conservation of the thermal current in the noninter-acting case, while, in the latter case, it is a consequence of afull suppression of any scattering between subspaces withdifferent numbers of doublons asU=t0 diverges. For a finite0 < U=t0 < ∞, 2πDth=I0 < 1, and it takes a minimumwith 2πDth=I0 ≈ 0.92 close to U ¼ 2t0, implying that atn ¼ 1, the dominant contribution to Re κðωÞ always comesfrom the Drude weight. Reference [9] provides a nonzerolower bound for Dth at T ¼ ∞ by considering only Q3

(a close relative of Ith [34]) in Eq. (2). By comparisonto this lower bound (solid line in the inset in Fig. 2),we conclude that the position of this minimum can beunderstood from the competition of the U-dependent andU-independent contributions to the Drude weight and to thetotal weight I0 ∝ hI2thi. Moreover, the lower bound fromRef. [9] is not exhaustive (see also Fig. S2 in Ref. [34]showing Dth and the lower bound as a function of n).Nonintegrable model and low-frequency dependence of

κregðωÞ.—Upon breaking integrability, our results for CthðtÞindicate a vanishing Drude weight, at least at high temper-atures and for intermediate values of U=t0. This raises thequestion of the functional form of κregðωÞ for V ≠ 0.Figure 3 shows κregðωÞ forU=t0¼ 4 at T ¼∞ and V¼U=3(main panel) and V ¼ 0 (inset). In the nonintegrable case,κregðωÞ has a broad peak at zero frequency, which is veryclose to a Lorentzian (the dashed line is a fit to the data).This demonstrates that standard diffusion is realized in theextended Hubbard model. In the integrable case, we oftenobserve maxima in κregðωÞ at ω > 0, which seem to berelated to the charge gap. Because of the uncertaintiesinvolved in extracting the frequency dependence, whichare due to the finite times reached in the simulations andthe extraction of the Drude weight, we are not able toresolve the low-frequency regime for V ¼ 0. Therefore, thequestion of whether κregðω → 0Þ is zero or finite in theintegrable case, which has been intensely studied for spintransport in the spin-1=2 XXZ chain [19,53,95–98],remains an open problem.

011T/t0

10−2

10−1

Dth

(T)/

t 02

U=0, exactU/t0=0U/t0=1U/t0=2U/t0=4U/t0=8

0 6 12U/t0

0.9

1

2 πD

th/I

0

T=∞

Dth(T=∞)/T2

Mazur bound

FIG. 2. Thermal Drude weight of the integrable model versusthe temperature for U=t0 ¼ 0, 1, 2, 4, 8. Solid line: Exact resultEq. (6) for U ¼ 0, in excellent agreement with DMRG data.Dashed lines: High-T behavior Dth ¼ D∞

th =T2, with D∞

th com-puted using DMRG. Inset: 2πDth=I0 at T ¼ ∞ (squares),compared to the lower Mazur bound [9] (solid line).

20 4ω /t0

0

2.4

4.8

T2κ r

eg(ω

)/t 0

3

DMRGLorentzian fit

0 4 8ω /t0

0

0.08

T2κ r

eg(ω

)/t 0

3

V=0

V=U/3

FIG. 3. Regular part κregðωÞ for U=t0 ¼ 4 at T ¼ ∞. Mainpanel: Nonintegrable case (V ¼ U=3). Dashed line: Fit of DMRGdata to a Lorentzian. Inset: Integrable case (V ¼ 0). The data areshown only for ω=t0 ≳ 1.4, whereas for smaller frequenciesuncertainties become too large (see the text).

PRL 117, 116401 (2016) P HY S I CA L R EV I EW LE T T ER Sweek ending

9 SEPTEMBER 2016

116401-3

Page 4: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

Spreading of local perturbations.—The presence of aballistic contribution in the linear response functionstranslates into the ballistic spreading of perturbations inthe local density [18,21,99–103]. To illustrate this con-nection, we study a Hubbard chain at an infinite temper-ature and introduce a perturbation in the local chargedensity at t ¼ 0þ. This also causes a perturbation in theenergy density. We measure the time evolution of bothdensities ρch;lðtÞ ¼ hnlðtÞi and ρth;lðtÞ ¼ hhlðtÞi, presentedin Figs. 4(a) and 4(b). While the energy density shows thetypical features of a ballistic dynamics [21,101], the chargedensity exhibits a much slower spreading and does not formfast ballistic jets, nicely illustrating the different nature ofenergy versus charge transport in this model. To becomemore quantitative, we compute the spatial variances asso-ciated with the density ρth;ch;lðtÞ:

σ2th;chðtÞ ¼1

N th;ch

XL−n0l¼n0

ðl − l0Þ2½ρth;ch;lðtÞ − ρbgth;ch�; ð7Þ

where l0 is the center of the wave packet, n0 cuts offboundary effects, ρbgth;ch;l denotes the bulk backgrounddensity, and N th;ch is the excess particle number or energyinduced by the wave packet. As expected, we find δσ2th ¼σ2thðtÞ − σ2thðt ¼ 0Þ ∝ t2 yet a much slower growth for thecharge σ2ch ∝ tα with 1=2 < α < 1 (see Fig. S4). Thedetermination of the exact exponent would require longertimes and is related to the low-frequency behavior of thecharge conductivity, yet, clearly, charge dynamics is notballistic. Another illustration for the ballistic energy

spreading can be obtained in T quenches [21], in whichwe embed a region with T2 into a larger system that is atT1 < T2, which overall has a homogeneous spin and chargedensity. An example is shown in Fig. 4(d), and, as expected,the variance of this wave packet grows as δσthðtÞ2 ∝ t2,illustrating that energy spreads ballistically in the temper-ature quench as well.The time evolution of the double occupancy (accessible

in optical-lattice experiments) dðtÞ ¼ hni↑ni↓iðtÞ is shownin Fig. 4(c) for the quench of Figs. 4(a) and 4(b). The profileexhibits fast ballistic jets, and the associated varianceσ2dðtÞ ¼ 1=D

Plðl − l0Þ2½hdlðtÞi − dbg� (D ¼ P

lhdli) in-creases approximately quadratically at long times and isthus sensitive to the fast spreading of the energy density. ForV ≠ 0 (see Fig. S5 for an illustration), the variances of bothenergy and double occupancy increase much slower thanquadratically in time. As a consequence of the ballisticenergy transport in a 1D FHM, we expect the absence ofthermalization in related quantum-gas experiments. From thelong-time behavior of the respective width δσth;chðtÞ, one canextract diffusion constants [18,21] or Drude weights (see,e.g., [103]) via Einstein relations (as we verified for ourcase), providing an experimental means of measuring trans-port coefficients. Such a local real-space and real-time probefor thermal transport has recently been used in experimentswith low-dimensional quantum magnets [104]. Given that acoupling to phonons cannot be avoided in quantum magnets[105,106], quantum-gas microscope experiments [58–64]could provide a means of studying energy and chargetransport in the FHM,which is easier to realizewith ultracoldquantum gases than the spin-1=2 XXZ chain in its massiveregime, where a similar coexistence of diffusive spin trans-port [20,21] and ballistic energy transport exists [9,11,107].Summary and outlook.—We computed the thermal

conductivity of the 1D FHM using a finite-T DMRGmethod. We confirm the ballistic nature of thermal trans-port in the integrable case, and we studied the temperaturedependence of the Drude weight. The lower bound for Dthfrom Ref. [9] is not exhaustive, implying that more local (oreven quasilocal [12,15,33]) conserved quantities than justQ3 play a role. We further demonstrated that the coexist-ence of diffusive charge transport and ballistic thermaltransport is directly reflected in local quantum quenchdynamics, presumably accessible to fermionic quantum-gas microscopes [58–64]. For the extended Hubbardmodel, we identified regimes in which, first, the Drudeweight clearly vanishes as system size increases and,second, the low-frequency dependence is compatible withdiffusive dynamics.From the theoretical point of view, an exact calculation

of κðωÞ exploiting the integrability of the model constitutesan open problem. In view of the existence of quasi-1Dmaterials described by the (extended) Hubbard model(including some Beechgard salts [108,109], organic mate-rials [110–112], and anorganic systems such as Sr2CuO3

FIG. 4. Local quenches for U=t0 ¼ 8 and V ¼ 0. (a)–(c) Attime t ¼ 0, a local excitation ðj0i þ j↑i þ j↓i þ 1.1j↑↓iÞ isprepared at two sites in the center of an otherwise equilibratedsystem at temperature T ¼ ∞. Real-time evolution of (a) thenormalized local charge density ½hnlðtÞi − 1�=n0, (b) the normal-ized local energy density hhlðtÞi=E0, and (c) the relativeprobability for double occupancy ½Pðnl ¼ 2Þ − 1=4�=p0. Thenormalization constants read n0¼ð2×1.12þ2Þ=ð1.12þ3Þ−1;E0≈0.025t0, and p0 ¼ 1.12=ð1.12 þ 3Þ − 1=4. (d) T quench(T1=t0¼10, T2¼∞): relative energy density hhlðtÞi=ð−0.245t0Þ.

PRL 117, 116401 (2016) P HY S I CA L R EV I EW LE T T ER Sweek ending

9 SEPTEMBER 2016

116401-4

Page 5: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

[113] and carbon nanotubes [114–116]), a detailed analysisof energy transport and the calculation of diffusion con-stants is desirable. Finally, the investigation of thermoelec-tric effects in SCSs is a timely topic (see, e.g., [45,117–121]) and should be feasible with our technique, at least athigh temperatures.

We thank C. Hess, T. Prosen, R. Steinigeweg, andX. Zotos for very helpful discussions. We thankM. Medenjak and T. Prosen for pointing out a mistakein a previous version of Fig. S2 to us. We acknowledgesupport by the Nanostructured Thermoelectrics program ofLBNL (C. K.) as well as by the DFG through the ResearchTraining Group 1995 (D. M. K.) and the Emmy Noetherprogram (C. K.).

[1] S. R. White, Phys. Rev. Lett. 69, 2863 (1992).[2] U. Schollwöck, Rev. Mod. Phys. 77, 259 (2005).[3] U. Schollwöck, Ann. Phys. (Berlin) 326, 96 (2011).[4] F. Essler, H. Frahm, F. Göhmann, A. Klümper, and

V. E. V. E. Korepin, The One-Dimensional Hubbard Model(Cambridge University Press, Cambridge, England, 2005).

[5] T. Giamarchi, Quantum Physics in One Dimension(Clarendon, Oxford, 2004), p. 2905.

[6] A. Klümper and D. C. Johnston, Phys. Rev. Lett. 84, 4701(2000).

[7] J.-S. Caux and J. M. Maillet, Phys. Rev. Lett. 95, 077201(2005).

[8] R. G. Pereira, J. Sirker, J.-S. Caux, R. Hagemans, J. M.Maillet, S. R. White, and I. Affleck, Phys. Rev. Lett. 96,257202 (2006).

[9] X. Zotos, F. Naef, and P. Prelovšek, Phys. Rev. B 55,11029 (1997).

[10] A. Klümper and K. Sakai, J. Phys. A 35, 2173 (2002).[11] K. Sakai and A. Klümper, J. Phys. A 36, 11617 (2003).[12] T. Prosen, Phys. Rev. Lett. 106, 217206 (2011).[13] J. Herbrych, P. Prelovšek, and X. Zotos, Phys. Rev. B 84,

155125 (2011).[14] C. Karrasch, J. Hauschild, S. Langer, and F. Heidrich-

Meisner, Phys. Rev. B 87, 245128 (2013).[15] T. Prosen and E. Ilievski, Phys. Rev. Lett. 111, 057203

(2013).[16] R. Steinigeweg, J. Gemmer, and W. Brenig, Phys. Rev.

Lett. 112, 120601 (2014).[17] P. Prelovšek, S. ElShawish, X. Zotos, and M.W. Long,

Phys. Rev. B 70, 205129 (2004).[18] R. Steinigeweg and J. Gemmer, Phys. Rev. B 80, 184402

(2009).[19] J. Sirker, R. G. Pereira, and I. Affleck, Phys. Rev. B 83,

035115 (2011).[20] R. Steinigeweg and W. Brenig, Phys. Rev. Lett. 107,

250602 (2011).[21] C. Karrasch, J. E. Moore, and F. Heidrich-Meisner, Phys.

Rev. B 89, 075139 (2014).[22] S. Kirchner, H. G. Evertz, and W. Hanke, Phys. Rev. B 59,

1825 (1999).[23] N. M. R. Peres, R. G. Dias, P. D. Sacramento, and J. M. P.

Carmelo, Phys. Rev. B 61, 5169 (2000).

[24] J. Carmelo, S.-J. Gu, and P. Sacramento, Ann. Phys.(Berlin) 339, 484 (2013).

[25] T. Prosen and M. Žnidarič, Phys. Rev. B 86, 125118 (2012).[26] T. Prosen, Phys. Rev. Lett. 112, 030603 (2014).[27] C. Karrasch, D. M. Kennes, and J. E. Moore, Phys. Rev. B

90, 155104 (2014).[28] F. Jin, R. Steinigeweg, F. Heidrich-Meisner, K. Michielsen,

and H. De Raedt, Phys. Rev. B 92, 205103 (2015).[29] G. D. Mahan, Many-Particle Physics (Plenum, New York,

1990).[30] J. M. Luttinger, Phys. Rev. 135, A1505 (1964).[31] M. Suzuki, Physica (Amsterdam) 51, 277 (1971).[32] P. Mazur, Physica (Amsterdam) 43, 533 (1969).[33] M. Mierzejewski, P. Prelovšek, and T. Prosen, Phys. Rev.

Lett. 114, 140601 (2015).[34] See Supplemental Material at http://link.aps.org/

supplemental/10.1103/PhysRevLett.117.116401 for defi-nitions, a discussion of transport coefficients, DMRG datafor several system sizes, a comparison of the thermalDrude weight of the FHM to the one of the Heisenbergchain, and additional results (Figs. S4 and S5) for localquenches, including Refs. [35–46].

[35] N. W. Ashcroft and N. D. Mermin, Solid State Physics(Saunders, Philadelphia, 1976).

[36] R. Steinigeweg (private communication).[37] D. Karevski and T. Platini, Phys. Rev. Lett. 102, 207207

(2009).[38] T. Prosen and B. Zunkovic, New J. Phys. 12, 025016 (2010).[39] M. Žnidarič, Phys. Rev. Lett. 106, 220601 (2011).[40] J. J. Mendoza-Arenas, T. Grujic, D. Jaksch, and S. R.

Clark, Phys. Rev. B 87, 235130 (2013).[41] A. V. Sologubenko, K. Gianno, H. R. Ott, U. Ammerahl,

and A. Revcolevschi, Phys. Rev. Lett. 84, 2714 (2000).[42] C. Hess, C. Baumann, U. Ammerahl, B. Büchner, F.

Heidrich-Meisner, W. Brenig, and A. Revcolevschi, Phys.Rev. B 64, 184305 (2001).

[43] A. E. Feiguin and G. A. Fiete, Phys. Rev. Lett. 106, 146401(2011).

[44] M. Soltanieh-ha and A. E. Feiguin, Phys. Rev. B 90,165145 (2014).

[45] M. R. Petersen, S. Mukerjee, B. S. Shastry, and J. O.Haerter, Phys. Rev. B 76, 125110 (2007).

[46] C. Hess, Dissertation, Universität zu Köln (2002)[47] C. Karrasch, J. Bardarson, and J. E. Moore, Phys. Rev.

Lett. 108, 227206 (2012).[48] C. Karrasch, J. H. Bardarson, and J. E. Moore, New J.

Phys. 15, 083031 (2013).[49] T. Barthel, New J. Phys. 15, 073010 (2013).[50] F. Verstraete, J. J. García-Ripoll, and J. I. Cirac, Phys. Rev.

Lett. 93, 207204 (2004).[51] A. E. Feiguin and S. R. White, Phys. Rev. B 72, 220401

(2005).[52] A. De Luca, J. Viti, L. Mazza, and D. Rossini, Phys. Rev. B

90, 161101 (2014).[53] C. Karrasch, D. M. Kennes, and F. Heidrich-Meisner,

Phys. Rev. B 91, 115130 (2015).[54] T. Barthel, U. Schollwöck, and S. R. White, Phys. Rev. B

79, 245101 (2009).[55] U. Schneider, L. Hackermüller, S. Will, T. Best, I. Bloch,

T. A. Costi, R. W. Helmes, D. Rasch, and A. Rosch,Science 322, 1520 (2008).

PRL 117, 116401 (2016) P HY S I CA L R EV I EW LE T T ER Sweek ending

9 SEPTEMBER 2016

116401-5

Page 6: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

[56] R. Jördens, N. Strohmaier, K. Günter, H. Moritz, and T.Esslinger, Nature (London) 455, 204 (2008).

[57] R. A. Hart, P. M. Duarte, T.-L. Yang, X. Liu, T. Paiva, E.Khatami, R. T. Scalettar, N. Trivedi, D. A. Huse, and R. G.Hulet, Nature (London) 519, 211 (2015).

[58] G. J. A. Edge, R. Anderson, D. Jervis, D. C. McKay, R.Day, S. Trotzky, and J. H. Thywissen, Phys. Rev. A 92,063406 (2015).

[59] A. Omran, M. Boll, T. A. Hilker, K. Kleinlein, G. Salomon,I. Bloch, and C. Gross, Phys. Rev. Lett. 115, 263001 (2015).

[60] E. Haller, J. Hudson, A. Kelly, D. Cotta, B. Peaudecerf, G.Bruce, and S. Kuhr, Nat. Phys. 11, 738 (2015).

[61] D. Greif, M. F. Parsons, A. Mazurenko, C. S. Chiu, S.Blatt, F. Huber, G. Ji, and M. Greiner, Science 351, 953(2016).

[62] M. Boll, T. A. Hilker, G. Salomon, A. Omran, I. Bloch, andC. Gross, arXiv:1605.05661.

[63] E. Cocchi, L. A. Miller, J. H. Drewes, M. Koschorreck, D.Pertot, F. Brennecke, and M. Köhl, Phys. Rev. Lett. 116,175301 (2016).

[64] L.W. Cheuk, M. A. Nichols, K. R. Lawrence, M. Okan, H.Zhang, and M.W. Zwierlein, Phys. Rev. Lett. 116, 235301(2016).

[65] C.-L. Hung, X. Zhang, N. Gemelke, and C. Chin, Phys.Rev. Lett. 104, 160403 (2010).

[66] B. Schmidt, M. R. Bakhtiari, I. Titvinidze, U. Schneider,M. Snoek, and W. Hofstetter, Phys. Rev. Lett. 110, 075302(2013).

[67] D. McKay and B. DeMarco, Rep. Prog. Phys. 74, 054401(2011).

[68] U. Schneider, L. Hackermüller, J. P. Ronzheimer, S. Will,S. Braun, T. Best, I. Bloch, E. Demler, S. Mandt, D. Rasch,and A. Rosch, Nat. Phys. 8, 213 (2012).

[69] S. Trotzky, Y.-A. Chen, A. Flesch, I. P. McCulloch, U.Schollwöck, J. Eisert, and I. Bloch, Nat. Phys. 8, 325(2012).

[70] J. Ronzheimer, M. Schreiber, S. Braun, S. Hodgman, S.Langer, I. McCulloch, F. Heidrich-Meisner, I. Bloch, andU. Schneider, Phys. Rev. Lett. 110, 205301 (2013).

[71] D. Pertot, A. Sheikhan, E. Cocchi, L. A. Miller, J. E. Bohn,M. Koschorreck, M. Köhl, and C. Kollath, Phys. Rev. Lett.113, 170403 (2014).

[72] G. Beni and C. F. Coll, Phys. Rev. B 11, 573 (1975).[73] R. Kubo, J. Phys. Soc. Jpn. 12, 570 (1957).[74] R. Kubo, M. Yokota, and S. Nakajima, J. Phys. Soc. Jpn.

12, 1203 (1957).[75] J. M. Luttinger, Phys. Rev. 135, A1505 (1964).[76] K. Louis and C. Gros, Phys. Rev. B 67, 224410 (2003).[77] J. Gemmer, R. Steinigeweg, and M. Michel, Phys. Rev. B

73, 104302 (2006).[78] M. Michel, G. Mahler, and J. Gemmer, Phys. Rev. Lett. 95,

180602 (2005).[79] T. Prosen and M. Žnidarič, J. Stat. Mech. (2009) P02035.[80] L. Arrachea, G. S. Lozano, and A. A. Aligia, Phys. Rev. B

80, 014425 (2009).[81] J. J. Mendoza-Arenas, S. Al-Assam, S. R. Clark, and D.

Jaksch, J. Stat. Mech. (2013) P07007.[82] M. Fannes, B. Nachtergaele, and R. F. Werner, J. Phys. A

24, L185 (1991).[83] S. Östlund and S. Rommer, Phys. Rev. Lett. 75, 3537 (1995).

[84] F. Verstraete and J. I. Cirac, Phys. Rev. B 73, 094423 (2006).[85] F. Verstraete, J. I. Cirac, and V. Murg, Adv. Phys. 57, 143

(2008).[86] F. Verstraete, J. J. Garcia-Ripoll, and J. I. Cirac, Phys. Rev.

Lett. 93, 207204 (2004).[87] S. White, Phys. Rev. Lett. 102, 190601 (2009).[88] M. Zwolak and G. Vidal, Phys. Rev. Lett. 93, 207205

(2004).[89] J. Sirker and A. Klümper, Phys. Rev. B 71, 241101(R)

(2005).[90] E. Jeckelmann, Phys. Rev. Lett. 89, 236401 (2002).[91] X. Zotos and P. Prelovšek, Phys. Rev. B 53, 983 (1996).[92] F. Heidrich-Meisner, A. Honecker, D. C. Cabra, and W.

Brenig, Phys. Rev. B 68, 134436 (2003).[93] X. Zotos, Phys. Rev. Lett. 92, 067202 (2004).[94] P. Jung, R. W. Helmes, and A. Rosch, Phys. Rev. Lett. 96,

067202 (2006).[95] J. Sirker, R. G. Pereira, and I. Affleck, Phys. Rev. Lett. 103,

216602 (2009).[96] S. Grossjohann and W. Brenig, Phys. Rev. B 81, 012404

(2010).[97] J. Herbrych, R. Steinigeweg, and P. Prelovšek, Phys. Rev.

B 86, 115106 (2012).[98] Y. Rahnavard, R. Steinigeweg, and W. Brenig, arXiv:

1501.03447.[99] M. Polini and G. Vignale, Phys. Rev. Lett. 98, 266403

(2007).[100] S. Langer, F. Heidrich-Meisner, J. Gemmer, I. McCulloch,

and U. Schollwöck, Phys. Rev. B 79, 214409 (2009).[101] S. Langer, M. Heyl, I. P. McCulloch, and F. Heidrich-

Meisner, Phys. Rev. B 84, 205115 (2011).[102] S. Jesenko and M. Žnidarič, Phys. Rev. B 84, 174438

(2011).[103] Y. Yan, F. Jiang, and H. Zhao, Eur. Phys. J. B 88, 11

(2015).[104] M. Otter, V. Krasnikov, D. Fishman, M. Pshenichnikov,

R. Saint-Martin, A. Revcolevschi, and P. van Loodsrecht,J. Magn. Magn. Mater. 321, 796 (2009).

[105] C. Hess, Eur. Phys. J. Spec. Top. 151, 73 (2007).[106] M. Montagnese, M. Otter, X. Zotos, D. A. Fishman, N.

Hlubek, O. Mityashkin, C. Hess, R. Saint-Martin, S.Singh, A. Revcolevschi, and P. H. M. van Loosdrecht,Phys. Rev. Lett. 110, 147206 (2013).

[107] F. Heidrich-Meisner, A. Honecker, D. C. Cabra, and W.Brenig, Phys. Rev. B 66, 140406 (2002).

[108] D. Jerome, Chem. Rev. 104, 5565 (2004).[109] V. Vescoli, L. Degiorgi, W. Henderson, G. Grüner, K. P.

Starkey, and L. K. Montgomery, Science 281, 1181 (1998).[110] T. Hasegawa, S. Kagoshima, T. Mochida, S. Sugiura, and

Y. Iwasa, Solid State Commun. 103, 489 (1997).[111] R. Claessen, M. Sing, U. Schwingenschlögl, P. Blaha, M.

Dressel, and C. S. Jacobsen, Phys. Rev. Lett. 88, 096402(2002).

[112] S. Wall, D. Brida, S. R. Clark, H. P. Ehrke, D. Jaksch,A. Ardavan, S. Bonora, H. Uemura, Y. Takahashi, T.Hasegawa, H. Okamoto, G. Cerullo, and A. Cavalleri,Nat. Phys. 7, 114 (2011).

[113] M. Ono, K. Miura, A. Maeda, H. Matsuzaki, H. Kishida,Y. Taguchi, Y. Tokura, M. Yamashita, and H. Okamoto,Phys. Rev. B 70, 085101 (2004).

PRL 117, 116401 (2016) P HY S I CA L R EV I EW LE T T ER Sweek ending

9 SEPTEMBER 2016

116401-6

Page 7: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

[114] M. Bockrath, D. H. Cobden, J. Lu, A. G. Rinzler, R. E.Smalley, L. Balents, and P. L. McEuen, Nature (London)397, 598 (1999).

[115] H. Ishii, H. Kataura, H. Shiozawa, H. Yoshioka, H.Otsubo, Y. Takayama, T. Miyahara, S. Suzuki, Y. Achiba,M. Nakatake, T. Narimura, M. Higashiguchi, K. Shimada,H. Namatame, and M. Taniguchi, Nature (London) 426,540 (2003).

[116] V. V. Deshpande, B. Chandra, R. Caldwell, D. Novikov,J. Hone, and M. Bockrath, Science 323, 106 (2009).

[117] B. Shastry, Rep. Prog. Phys. 72, 016501 (2009).[118] P. Wissgott, A. Toschi, H. Usui, K. Kuroki, and K. Held,

Phys. Rev. B 82, 201106 (2010).[119] S. Hong, P. Ghaemi, J. E. Moore, and P.W. Phillips, Phys.

Rev. B 88, 075118 (2013).[120] M. Mierzejewski, D. Crivelli, and P. Prelovšek, Phys. Rev.

B 90, 075124 (2014).[121] J. Kokalj and R. H. McKenzie, Phys. Rev. B 91, 125143

(2015).

PRL 117, 116401 (2016) P HY S I CA L R EV I EW LE T T ER Sweek ending

9 SEPTEMBER 2016

116401-7

Page 8: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

Thermal conductivity of the one-dimensional Fermi-Hubbard model:Supplemental Material

C. Karrasch,1, 2 D. M. Kennes,3 and F. Heidrich-Meisner4

1Department of Physics, University of California, Berkeley, California 95720, USA2Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA

3Institut fur Theorie der Statistischen Physik, RWTH Aachen University andJARA-Fundamentals of Future Information Technology, 52056 Aachen, Germany4Department of Physics and Arnold Sommerfeld Center for Theoretical Physics,

Ludwig-Maximilians-Universitat Munchen, 80333 Munchen, Germany

0 2 4t t0

0

4

8

Cth

(t)/

t 04

L=10L=16L=20L=40L=100

0 2 4

0

0.2

T/t0=2

T/t0=0.66

T=∞

FIG. S1. (Color online) Energy-current autocorrelation func-tion at U/t0 = 8 and V = 0 calculated for various systemsizes L. Note that the curves for L = 40 and L = 100 areindistinguishable.

S1. Energy current operator in fermionic and spinlanguage

Following Ref. 1, we define the energy current op-erator using a continuity equation. This yields Ith =i∑L−2l=1 [hl, hl+1]. For V = 0, this results in

Ith =∑l,σ

t20

[(ic†l+1σcl−1σ + h.c.

)− U

2

(jl−1σ + jlσ

)(nlσ −

1

2

)],

(S1)

where jlσ = it0c†l+1σclσ + h.c. is the local charge current.

Ith has the same structure as the conserved charge Q3,except for a different prefactor of 1/2 in front of the U -dependent term [1].

Within the DMRG numerics, we implement a spin ver-sion of the Hamiltonian as well as of Ith, which we obtainvia a Jordan-Wigner transformation:

cl↑ = (σz1 · · ·σzl−1)σ−l , cl↓ =( L∏l=1

σzl

)(τz1 · · · τzl−1)τ−l ,

(S2)where σx,y,zl and τx,y,zl denote standard Pauli matrices,and σ±l = (σxl ± iσ

yl )/2, τ±l = (τxl ± iτ

yl )/2. This leads

0 0.2 0.4 0.6 0.8 1filling

0

0.2

0.4

0.6

T2D

th(T

)/t 0

4

DMRGMazur bound

FIG. S2. (Color online) Filling-dependence of the Drudeweight at U/t0 = 2 and T = ∞. The lower Mazur boundfrom Ref. 1 is shown as a comparison.

to (for notational simplicity we again set V = 0)

H =∑l

[t0(σ+l+1σ

−l + τ+

l+1τ−l + h.c.

)+U

4σzl τ

zl

], (S3)

and

Ith =∑l

{(− it20σ+

l+1σzl σ−l−1 − it

20τ

+l+1τ

zl τ−l−1 + h.c.

)− U

4

[(jσl−1 + jσl

)τzl +

(jτl−1 + jτl

)σzl

]},

(S4)

where jσl = −it0σ+l+1σ

−l + h.c., jτl = −it0τ+

l+1τ−l + h.c..

S2. Transport coefficients

In general, the heat current Iq of an electronic systemis defined as [2, 3]

Iq = Ith − µI . (S5)

As a consequence, one needs to consider a two-by-twomatrix of transport coefficients Lij that relate externalforces to currents. One possible choice is to work with theheat current Iq and the particle(electrical) current I and

Page 9: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

2

0.1 1 10T/t0

0.001

0.01

0.1

1

Dth

(T)/t

02

Heisenberg

Hubbard

FIG. S3. (Color online) Drude weight of the Hubbard modelat large U/t0 = 8 in comparison with the result for anisotropic Heisenberg spin chain with an antiferromagneticcoupling J = 4t20/U .

0 2 4 6t t0

0

40

80

δσ

2 ν(t

)

energy

charge

doubleocc.

FIG. S4. (Color online) Variances δσ2th,ch,d(t) for the data

shown in Figs. 4(a-c) of the main text. The dashed linesshow fits to δσ2

ν(t) ∼ tα (ν = th,ch,d) for times tt0 > 3. Thefitted exponents read αth = 1.96, αch = 1.28, and αd = 1.76.

the corresponding forces FJ = − 1T∇(µ + V ), where V

is the electrostatic potential, µ is the chemical potentialand Fq = ∇(1/T ), resulting in(

IIq

)=

(L11 L12

L21 L22

)(FJFq

)(S6)

(we set the electrical charge to one for simplicity). Thetransport coefficients Lij are computed from the respec-tive Kubo formulae [2].

The thermal conductivity κ is measured under the con-dition of 〈I〉 = 0 and, therefore, is given by

κ = L22 − L21L−111 L12 . (S7)

Alternatively, one can work with the currents Ith and Iand corresponding transport coefficients Mij . The ther-mal conductivity is then given by

κ = M22 −M21M−111 M12 . (S8)

Thus, in principle, two terms contribute to κ, the secondarises because of the thermoelectric coupling. As is dis-cussed in textbooks, Eqs. (S7) and (S8) yield equivalentresults for κ.

Let us first consider the corresponding Drude weightsDij associated with the Mij (in the notation of the maintext, D11 = Ds and D22 = Dth). In the case of ahalf-filled system, e.g., the Mott insulator in 1D thatis considered in the main text, the energy and parti-cle current have a different parity under particle-holesymmetry and therefore, the current-correlation functionCI,Ith(t) = 〈I(t)Ith〉 vanishes identically. As a conse-quence, M12 = M21 = 0 at half filling at all temper-atures. This result is known from studies of, e.g., thethermopower which consequently also vanishes in a Mottinsulator at n = 1 in the presence of particle-hole sym-metry [4, 5]. In other words, since we wrote down theHamiltonian (3) in an explicitly particle-hole symmetricform for V = 0, the chemical potential vanishes at halffilling. For the purpose of our work we conclude that theDrude weight Dq associated with the heat current Iq isgiven by

Dq = Dth . (S9)

Therefore, for the case studied in the main text,namely half filling, the thermal conductivity solely stemsfrom the energy-current correlation function Cth(t) =〈Ith(t)Ith〉.

Note that the justification of the Kubo formula for thethermal conductivity [6] is more subtle than for chargetransport (see [7] for a comprehensive discussion and ad-ditional references). The main reason is that a gradient intemperature is not a mechanical force in the same senseas a voltage or gradient in chemical potential is. There-fore, one cannot add an additional term to the Hamilto-nian that would contain the temperature gradient, whichone would commonly treat as the perturbation drivingthe system out of equilibrium. An approach to circum-vent that problem is to introduce so-called polarizationoperators (see, e.g., [8] for a discussion). Such polariza-tion operators (essentially of the form P = −

∑l lhl) can

then take the role of the external force. This construc-tion, however, requires open boundary conditions. Otherattempts make an assumption about local equilibrium orrely on the so-called entropy production argument (seeagain [7] for details).

Ref. [7] provides a derivation of the Kubo formula thatdoes not rely on the assumption of local equilibrium.That paper also attempts a direct comparison betweentime-dependent numerical solutions of certain models inthe presence of a temperature gradient and mostly con-firms the validity of the Kubo formula. Some discrepan-cies were observed in Ref. [7], which, however, were laterunderstood to be finite-size effects [9].

There is also substantial interest in extracting ther-mal currents and temperature profiles from simulationsof open quantum systems (see, e.g., [10–17]), where typi-cally, master equations are solved. In some of these stud-ies (see, e.g., [15]), however, transport is boundary-driven,which realizes a different physical situation from what

Page 10: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

3

underlies the derivation of Kuba formulae for charge,spin and energy transport, where the bulk experiencesa gradient of some force. As a consequence, such openquantum system simulations can yield qualitatively dif-ferent results compared to the Kubo formula, even forcharge or spin transport (compare, e.g., [15, 18]). More-over, it is important to keep in mind that conductivitiesare bulk properties by definition. Thus, any contact re-gions must be excluded in measuring currents and, inparticular, temperature differences. Therefore, experi-ments commonly use a four-terminal setup to measurethe thermal conductivity (see, e.g., [19–21]). The techni-cal challenge for such open quantum system simulationsis therefore to be able to reach sufficiently large systems,as a result of which a direct comparison to the Kuboformula is often not possible.

S3. Comparison of DMRG results for differentsystem sizes

In Fig. S1, we present DMRG results for Cth(t) for sev-eral different system sizes and temperatures at U/t0 = 8and V = 0. Since we are using open boundary conditions,finite-size effects manifest themselves by a drop of Cth(t)to zero and a sign change at a system-size dependent timet∗(L). Very importantly, for t < t∗(L), the data from dif-ferent system sizes all coincide, implying that the resultsare for the thermodynamic limit. Moreover, using datafor open boundary conditions, one needs at least L & 20to get to times of the order of 4 ∼ 1/t0. Using L = 100,we clearly never reach t∗, hence the time-dependent datafrom such a large system is free of finite-size effects in thetime window reached in the simulations. Note that thenumerical effort is essentially independent of L but pri-marily depends on the entanglement growth that limitsthe accessible times.

S4. Comparison of the Drude weights of the FHMand Heisenberg chain

In principle, two types of excitations should contributeto the thermal conductivity. First, for temperaturesT > ∆Mott, where ∆Mott is the Mott gap, charge ex-citations (i.e., excitations in the upper Hubbard band)become relevant. Second, at very low T � ∆Mott, chargeexcitations are frozen out and then gapless spin excita-tions should be the only available ones. Whether thesetwo regimes can be resolved depends on the ratio of thecharacteristic energy scale J for magnetic excitations andthe Mott gap ∆Mott. In the large U/t0 limit, J = 4t20/U ,much smaller than ∆Mott ∼ U . Therefore, the very lowtemperature dependence of Dth of the Hubbard modelshould be identical to the one of the spin-1/2 Heisen-berg model [22, 23], where Dth increases linearly at low

FIG. S5. (Color online) The same as in Figs. 4(a-c) ofthe main text, but for V = U/4. The exponents of thetime-dependence of the corresponding width δσ2

ν ∝ tα (ν =th,ch,d) are αth = 1.14, αch = 0.72, and αd = 0.99.

T , takes a maximum at T . J/2 and then decreases tozero with 1/T 2. As a consequence, the full Dth for largeU/t0 could have a double maximum structure and more-over, at T � J , the conductivity should solely stem fromcharge excitations since the spin system will effectivelybe at T/J = ∞ (This is often referred to as the spin-incoherent regime, see, e.g., [24, 25]). The Drude weightsof the FHM at U/t0 = 8 and the spin-1/2 Heisenbergchain [22] are shown in Fig. S3, confirming the qualita-tive picture described above. Note also the difference inthe maximum values of the Drude weights: the chargedominated Dth exceeds the Heisenberg contribution sig-nificantly. The problem with numerically resolving theinteresting crossover regime is that it requires a large U ,which in the units of the FHM renders J a very smalltemperature that at present cannot be reached.

[1] X. Zotos, F. Naef, and P. Prelovsek, Phys. Rev. B 55,11029 (1997).

Page 11: Thermal Conductivity of the One-Dimensional Fermi-Hubbard Model

4

[2] G. D. Mahan, Many-Particle Physics (Plenum Press,New York London, 1990).

[3] N. W. Ashcroft and N. D. Mermin, Solid State Physics(Saunders College Publishing, 1976).

[4] G. Beni and C. F. Coll, Phys. Rev. B 11, 573 (1975).[5] M. R. Peterson, S. Mukerjee, B. S. Shastry, and J. O.

Haerter, Phys. Rev. B 76, 125110 (2007).[6] J. M. Luttinger, Phys. Rev. 135, A1505 (1964).[7] J. Gemmer, R. Steinigeweg, and M. Michel, Phys. Rev.

B 73, 104302 (2006).[8] K. Louis and C. Gros, Phys. Rev. B 67, 224410 (2003).[9] R. Steinigeweg, , private communication.

[10] M. Michel, G. Mahler, and J. Gemmer, Phys. Rev. Lett.95, 180602 (2005).

[11] T. Prosen and M. Znidaric, J. Stat. Mech 2009, P02035(2009).

[12] D. Karevski and T. Platini, Phys. Rev. Lett. 102, 207207(2009).

[13] L. Arrachea, G. S. Lozano, and A. A. Aligia, Phys. Rev.B 80, 014425 (2009).

[14] T. Prosen and B. Zunkovic, New J. Phys. 12, 025016(2010).

[15] M. Znidaric, Phys. Rev. Lett. 106, 220601 (2011).[16] J. J. Mendoza-Arenas, S. Al-Assam, S. R. Clark, and

D. Jaksch, J. Stat. Mech. , P07007 (2013).[17] J. J. Mendoza-Arenas, T. Grujic, D. Jaksch, and S. R.

Clark, Phys. Rev. B 87, 235130 (2013).[18] C. Karrasch, J. E. Moore, and F. Heidrich-Meisner,

Phys. Rev. B 89, 075139 (2014).[19] A. V. Sologubenko, K. Gianno, H. R. Ott, U. Ammerahl,

and A. Revcolevschi, Phys. Rev. Lett. 84, 2714 (2000).[20] C. Hess, C. Baumann, U. Ammerahl, B. Buchner,

F. Heidrich-Meisner, W. Brenig, and A. Revcolevschi,Phys. Rev. B 64, 184305 (2001).

[21] C. Hess, (Dissertation, Universitat zu Koln, 2002).[22] A. Klumper and K. Sakai, J. Phys. A 35, 2173 (2002).[23] F. Heidrich-Meisner, A. Honecker, D. C. Cabra, and

W. Brenig, Phys. Rev. B 66, 140406 (2002).[24] A. E. Feiguin and G. A. Fiete, Phys. Rev. Lett. 106,

146401 (2011).[25] M. Soltanieh-ha and A. E. Feiguin, Phys. Rev. B 90,

165145 (2014).