This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Resolving the Thermal Paradox: Many-body localization or fractionalization?

Saikat Banerjee saikat.banerjee@rutgers.edu Center for Materials Theory, Rutgers University, Piscataway, New Jersey, 08854, USA    Piers Coleman pcoleman@physics.rutgers.edu Center for Materials Theory, Rutgers University, Piscataway, New Jersey, 08854, USA Hubbard Theory Consortium, Department of Physics, Royal Holloway, University of London, Egham, Surrey TW20 0EX, UK
Abstract

Thermal measurements of heat capacity and thermal conductivity in a wide range of insulators and superconductors exhibit a “thermal paradox”: a large linear specific heat reminiscent of neutral Fermi surfaces in samples that exhibit no corresponding linear temperature coefficient to the thermal conductivity. At first sight, these observations appear to support the formation of a continuum of thermally localized many-body excitations, a form of many-body localization that would be fascinating in its own right. Here, by mapping thermal conductivity measurements onto thermal RC circuits, we argue that the development of extremely long thermal relaxation times, a “thermal bottleneck,” is likely in systems with either many-body localization or neutral Fermi surfaces due to the large ratio between the electron and phonon specific heat capacities. We present a re-evaluation of thermal conductivity measurements in materials exhibiting a thermal paradox that can be used in future experiments to deliberate between these two exciting alternatives.

Introduction. Several fascinating non-metallic quantum materials (see Table  1) exhibit large linear specific heat capacities CV=γTC_{V}=\gamma T, a hallmark of a neutral Fermi surface [1], yet show no corresponding linear coefficient in the thermal conductivity, k0=κ/Tk_{0}=\kappa/T, at low temperatures. From the Einstein relation κ=DCV\kappa=DC_{V} where DD is the particulate diffusion constant, CVC_{V} is the specific heat capacity 111The corresponding Einstein relation for the conductivity of a metal is responsible for the Weidemann-Franz ratio between thermal κ\kappa and electrical conductivity σ\sigma, κ/T=(π2kB2/3e2)σ{\kappa}/{T}=(\pi^{2}k_{B}^{2}/3e^{2})\sigma we expect [3]

k0=Dγ.k_{0}=D\gamma. (1)

Thus, at first sight, the absence of a thermal conductivity in a system with a linear specific heat suggests the absence of diffusion D=0D=0 in a neutral quasiparticle continuum: a form of many-body localization [4]. In this letter, we identify an important thermal bottleneck which can however obscure any underlying quasiparticle thermal conductivity due to long equilibration times between the phonons and neutral quasiparticles. Our main conclusion is that further measurements, keeping careful track of equilibration times are required to discern between thermal localization and neutral Fermi surfaces.

Examples of systems exhibiting a thermal paradox are given in Table 1). For instance, the Kondo insulator, Samarium Hexaboride (SmB6) has a linear specific heat in the range γ550\gamma\sim 5-50 mJK-2mol-1 [5], considerably larger than the isostructural metal Lanthanum Hexaboride (LaB6), but it exhibits a vanishing residual thermal conductivity at zero magnetic field [6]. In this particular case, previous experiments have also reported evidence for quantum oscillations in magnetic torque [7, 8] and specific heat measurements [9] that have been interpreted in terms of excitations above neutral Fermi surfaces with long mean-free paths. Similar paradoxes arise in spin-liquid candidates such as the organic insulator β\beta^{\prime}-EtMe3Sb[Pd(dmit)2]2 (dmit-131), which exhibits the absence of a linear thermal conductivity coefficient despite a sizeable linear specific heat [10, 11]. The case of the heavy fermion superconductor Uranium diTelluride (UTe2) is another interesting example. In this case, samples prepared by chemical vapor deposition display a linear specific heat γ60\gamma\sim 60 mJK-2mol-1 in the superconducting state, one-half the normal state value [12, 13]. While the linear specific heat disappears in higher-quality flux-grown samples, it is still surprising that samples with half the normal state density of states exhibit no thermal diffusion.

System Insulator Quantum spin liquid Superconductor
SmB6 YbB12 dmit-131 NaYbS2, NaYbSe2 1T-TaS2 UTe2
Ref [5, 6, 14] [15] [16, 11, 17] [18, 19, 20] [21, 22] [23, 12, 13, 24]
γ\gamma (mJK-2 mol-1) 25\sim 25 3.8\sim 3.8 20\sim 20 1.1\sim 1.1 1.84\sim 1.84 60\sim 60
β\beta (mJK-4mol-1) 0.45\sim 0.45 0.026\sim 0.026 24\sim 24 0.1\sim 0.1 0.31\sim 0.31 2.81\sim 2.81
Cf/Cp=γ/βT02C_{f}/C_{p}=\gamma/\beta T_{0}^{2} 350\sim 350 584\sim 584 40\sim 40 1000\sim 1000 590\sim 590 133\sim 133
Base temperature T0T_{0} (K) 0.4\sim 0.4 0.5\sim 0.5 0.15\sim 0.15 0.1\sim 0.1 0.1\sim 0.1 0.4\sim 0.4
κ0/T\kappa_{0}/T (mWK-2cm-1) 0\sim 0 0.01\sim 0.01 0\sim 0 0\sim 0 0\sim 0 0\sim 0
Table 1: Measured values of the various coefficients in the specific heat capacity C/T=γ+βT2C/T=\gamma+\beta T^{2}, (γ\gamma is typically attributed to spin/fermions, and β\beta to phonons) and residual thermal conductivity κ0/T\kappa_{0}/T at low temperatures. Cf/Cp=γ/βT02C_{f}/C_{p}=\gamma/\beta T_{0}^{2} gives a ratio of the spin/fermion to phonon specific heat at the base measurement temperature T0T_{0}. (The phonon contribution is measured from the isostructural compound NaYbO2 [25]).

Yamashita et al. [11] – have recently noted that the measured low-temperature thermal conductivity depends on the cooling rate used in the experiment. This suggests that thermal equilibrium in the temperature gradient is not established without a very slow measurement protocol. This conclusion is corroborated by more recent measurements on dmit-131 [17], which hint at a subtle spin-phonon coupling responsible for the slow transmission of heat into the underlying spin fluid.

To investigate thermal relaxation rates from various possible sources, we map the problem onto a thermal circuit diagram [26]. The sample is modeled as a one-dimensional object with endpoints at two distinct temperatures. The steady-state heat thermal current between its two ends, ii and jj, is described by the thermal Ohm’s law

IT(ij)=(κAl)(TiTj)I_{T}(i\rightarrow j)=\left(\frac{\kappa A}{l}\right)(T_{i}-T_{j}) (2)

where ITI_{T} is the heat current flowing from ii to jj, TiTjT_{i}-T_{j} is the difference in temperatures, κ\kappa is the thermal conductivity, AA and ll are the cross-sectional area and length of the sample, respectively and κ\kappa is the total thermal conductivity, i.e., an aggregate of contributions from various channels such as phonons, electrons and other quasiparticles κ=κp+κe+\kappa=\kappa_{\rm{p}}+\kappa_{e}+\ldots. The conservation of heat in a thermal circuit allows us to exploit the analogy between heat and charge QTQQ_{T}\leftrightarrow Q, thermal and electrical current ITIeI_{T}\leftrightarrow I_{e}. We can identify temperature with voltage (TiViT_{i}\leftrightarrow V_{i}) and Ri=l/(κiA)R_{i}=l/(\kappa_{i}A) as the thermal resistance of a given thermal conduction channel. In a similar fashion, we can identify a heat bath with a capacitor in which the heating rate is equal to the incoming thermal current dQT/dt=ITdQ_{T}/dt=I_{T}: by comparing dQT/dT=CTdQ_{T}/dT=C_{T} with dQ/dV=CedQ/dV=C_{e}, we may identify the specific heat CT=CeC_{T}=C_{e} as a thermal capacitance. These connections allow us to represent heat conduction circuits as RC electrical circuits [27, 28, 29], allowing us to analyze various thermal relaxation timescales in terms of the corresponding time constants τCTRT\tau\sim C_{T}R_{T}, as summarized in Table 2. Curiously, thermal RC circuits have not been widely used in a condensed matter framework, but they have been extensively used to model heat transfer in urban buildings and climate sciences [30, 31].

Electrical Thermal
Quantity Q=Q= Charge QT=Q_{T}= Heat
Potential VV (voltage) TT (temperature)
Ohm’s Law V(ω)=I(ω)Z(ω)V(\omega)=I(\omega)Z(\omega) T(s)=IT(s)Z(s)T(s)=I_{T}(s)Z(s)
ZRZ_{R} ReR_{e} RT=lAκR_{T}=\frac{l}{A\kappa}
ZCZ_{C} 1iωC\frac{1}{-i\omega C} 1sC\frac{1}{-sC}
Table 2: Mapping between electrical and thermal circuits.

Model. We consider the effect of weak thermal coupling between the phonon and spin/fermion degrees of freedom on the equilibration times in a specific heat and a thermal conductivity measurement, introducing two circuit models, depicted in Fig. 1(a,b). Fig. 1(a) illustrates the circuit for a specific heat measurement. In a specific heat measurement, a pulse of heat is sent into the thermally isolated system. The heat initially flows into the phonon modes of the material, represented by capacitance CpC_{\rm p}. As the lattice temperature rises, heat is transferred into the electronic and magnetic modes of the material via the thermal resistance RpfR_{\rm pf}, flowing into the spin/fermion degrees of freedom, represented by capacitance CfC_{\rm f}. Fig. 1(b) shows the corresponding circuit for thermal conductivity measurement. In this case, a constant heat flux is applied to one surface, measuring the steady-state temperature gradient once equilibrium is established. The heat initially accumulates in the phonon modes, represented by CpC_{\rm{p}}, and is then transmitted into the electronic and magnetic modes via the thermal resistance RpR_{\rm{p}}. The spin/fermion degrees of freedom conduct heat via a thermal resistance RfR_{\rm{f}} which discharges the heat in the spin-fermion system carried by thermal capacitance CfC_{\rm{f}}. If many-body localized spin/fermion degrees of freedom are present, RfR_{\rm{f}} is essentially infinite. Although these models provide a simplified representation of the measurement protocols, they capture the key phenomenological features of thermal relaxation.

Refer to caption
Figure 1: Thermal circuit diagrams describing (a) specific heat and (b) thermal conductivity measurements.

In a canonical insulator, the low-temperature phonon specific heat is a cubic function of temperature; we are particularly interested in the case where the electronic/spin degrees of freedom exhibit a linear specific heat so that

Cf=γT,Cp=βT3.C_{\rm{f}}=\gamma T,\quad C_{\rm{p}}=\beta T^{3}. (3)

Experimental estimates of β,γ\beta,\gamma for various materials are provided in Table 1.

Impedance and temperature profile. Table 1 summarizes the analogy between electrical and thermal circuits. Since thermal equilibration is most conveniently described in terms of Laplace transforms, we write the temperature as transformed as follows

T(s)=0T(t)est𝑑t.T(s)=\int_{0}^{\infty}T(t)e^{-st}dt.

In the Laplace transform ss plays the role of an imaginary frequency, and in the generalized thermal Ohm’s law T(s)=IT(s)Z(s)T(s)=I_{T}(s)Z(s) we must then replace iωsi\omega\rightarrow s in each of the circuit elements, so that a heat bath, as a thermal capacitor is associated with an impedance Z(s)=1sCZ(s)=\frac{1}{-sC}.

We now compute the impedances for each of the circuits depicted in Fig. 1(a,b). A key advantage of this analysis is that the location of the poles sl=iΓls_{l}=i\Gamma_{l} determines the thermal equilibration rates Γl=1/τl\Gamma_{l}=1/\tau_{l} of the circuit. For specific heat measurements, the thermal impedance (see Supplementary Material (SM) [32] for details) involves a single pole with the time constant as

1τ0=1Rpf(1Cp+1Cf).\frac{1}{\tau_{0}}=\frac{1}{R_{\rm{pf}}}\left(\frac{1}{C_{\rm{p}}}+\frac{1}{C_{\rm{f}}}\right). (4)

In the limit of large spin/fermion capacitance (γβT02\gamma\gg\beta T_{0}^{2}, where T0T_{0} is the characteristic temperature scale for the measurement, see Table 1), the time-constant becomes τ0=RpfCp\tau_{0}=R_{\rm{pf}}C_{\rm{p}}, indicating that the large thermal resistances of the spin fluid and phonon do not affect the time required for heat capacity measurement.

However, in a thermal conductivity measurement, modeled by the circuit diagram in Fig. 1(b), the thermal resistances for both the spin/fermion and phonon play a role in the equilibration, leading to fast and a slow time constants

τF=CpRpfRpRpf+Rp,τS=CfRf(Rp+Rpf)Rpf+Rp+Rf.\tau_{\rm{F}}=\frac{C_{\rm{p}}R_{\rm{pf}}R_{\rm{p}}}{R_{\rm{pf}}+R_{\rm{p}}},\quad\tau_{\rm{S}}=\frac{C_{\rm{f}}R_{\rm{f}}(R_{\rm{p}}+R_{\rm{pf}})}{R_{\rm{pf}}+R_{\rm{p}}+R_{\rm{f}}}. (5)

It is the large ratio between the specific heat of the electron and phonon fluids that sets the ratio between τS/τFCf/Cp1\tau_{\rm S}/\tau_{\rm F}\sim C_{\rm f}/C_{\rm p}\gg 1. Whereas the fast time-constant τF\tau_{\rm{F}} is even shorter than the equilibration time of a specific heat measurement, the slow time-constant τS\tau_{\rm{S}} is much larger than both τ0\tau_{0} and τF\tau_{\rm{F}}. We can understand these very different time scales by noting that at the short times governing τS\tau_{\rm S}, the thermal capacitance CfC_{\rm{f}} is shorted out, while at the long-times governing τS\tau_{S}, CfC_{\rm{f}} is in series with the parallel resistors Rpf+RpR_{\rm{pf}}+R_{\rm{p}}, and RfR_{\rm{f}}. In the limit of large spin/fermion thermal resistance (RfR_{\rm{f}}\to\infty), the fast time-constant τF\tau_{\rm{F}} remains the same, while the slow time-constant τS\tau_{\rm{S}} becomes much longer. Note that within our model RfR_{\rm{f}}\to\infty would serve as a potential signature of many-body localization.

The temperature profile as a function of time in each of these circuits is obtained by inverting the Laplace transform T(s)=I(s)Z(s)T(s)=I(s)Z(s) back into the time domain, which yields

T(t)=I0[A1(1et/τF)+A2(1et/τS)],T(t)=I_{0}\biggl{[}A_{1}(1-e^{-t/\tau_{\rm{F}}})+A_{2}(1-e^{-t/\tau_{\rm{S}}})\biggr{]}, (6)

where A1,A2A_{1},A_{2} are constants depending on the thermal resistances and capacitances defined in Fig. 1(b), and I0I_{0} is the long-time equilibration thermal current. A detailed derivation can be found in SM [32]. The corresponding temperature evolution relevant for the thermal conductivity measurement is shown in Fig. 2 for a very large spin fluid capacitance (CfCpC_{\rm{f}}\gg C_{\rm{p}}), and assuming almost identical thermal resistances (RpRpfRfR_{\rm{p}}\sim R_{\rm{pf}}\sim R_{\rm{f}}). It is clear from Fig. 2 that the large capacitance CfC_{\rm{f}} only affects the time constant for measuring thermal conductivity, not the time constant for measuring the specific heat.

Refer to caption
Figure 2: An illustration of the time-dependent temperature profile in case of thermal conductivity measurement, obtained by utilizing Eq. (6). For quantitative purposes, we adopt Cp=1C_{\rm{p}}=1, Cf=1000C_{\rm{f}}=1000, and Rp=Rpf=Rf=1R_{\rm{p}}=R_{\rm{pf}}=R_{\rm{f}}=1 in arbitrary unit. The ratio Cf:CpC_{\rm{f}}:C_{\rm{p}} is chosen as a large value motivated by the experimental estimates for specific heat capacity (γ\gamma, and β\beta) in various materials, see Table. 1.

In a steady-state measurement, thermal capacitances can be neglected under the assumption that they are fully charged. Under this condition, we can estimate the secondary temperature increment predicted in Fig. 2 using the circuit representation in Fig. 1(a). The temperature difference is determined from the thermal resistances as

1ΔTF1ΔTI=1I0(1Rpf1Rpf+Rf),\frac{1}{\Delta T_{\rm{F}}}-\frac{1}{\Delta T_{\rm{I}}}=\frac{1}{I_{0}}\left(\frac{1}{R_{\rm{pf}}}-\frac{1}{R_{\rm{pf}}+R_{\rm{f}}}\right), (7)

where ΔTI\Delta T_{\rm{I}} (ΔTF\Delta T_{\rm{F}}) is the initial (final) temperature increment as visualized in Fig. 2. This simple result shows the necessity of long waiting times to determine whether the underlying system contains neutral quasiparticles or many-body localized excitations.

Discussion and conclusion. The paradox of coexisting linear specific heat with negligible residual thermal conductivity in various quantum materials signals a need for a deeper understanding of heat transport mechanisms in systems with neutral fermionic excitations. While the lack of thermal diffusion may initially suggest evidence of many-body localization, our analysis suggests that slow thermal equilibration – primarily due to weak phonon-quasiparticle coupling – could serve as an alternative explanation.

By modeling thermal transport in these systems using an electrical circuit analogy, we have identified key relaxation timescales that may obscure the intrinsic heat conduction properties of the spin and fermion degrees of freedom. Our findings indicate that it is crucial to reconsider current measurement protocols, particularly in relation to equilibration times, to distinguish between true thermal localization and hidden neutral Fermi surfaces. Further experimental investigations, including time-resolved thermal transport measurements, are necessary to fully understand the nature of these low-temperature excitations across these class of materials.

Acknowledgments. This work is supported by the Office of Basic Energy Sciences, Material Sciences and Engineering Division, U.S. Department of Energy (DOE) under Contracts No. DE-FG02-99ER45790.

References

  • Baym and Pethick [2008] G. Baym and C. Pethick, Landau Fermi-Liquid Theory: Concepts and Applications (Wiley, 2008).
  • Note [1] The corresponding Einstein relation for the conductivity of a metal is responsible for the Weidemann-Franz ratio between thermal κ\kappa and electrical conductivity σ\sigma, κ/T=(π2kB2/3e2)σ{\kappa}/{T}=(\pi^{2}k_{B}^{2}/3e^{2})\sigma.
  • Bird et al. [2002] R. Bird, W. Stewart, and E. Lightfoot, Transport Phenomena (J. Wiley, 2002).
  • Abanin et al. [2019] D. A. Abanin, E. Altman, I. Bloch, and M. Serbyn, Colloquium: Many-body localization, thermalization, and entanglement, Rev. Mod. Phys. 91, 021001 (2019).
  • Phelan et al. [2014] W. A. Phelan, S. M. Koohpayeh, P. Cottingham, J. W. Freeland, J. C. Leiner, C. L. Broholm, and T. M. McQueen, Correlation between Bulk Thermodynamic Measurements and the Low-Temperature-Resistance Plateau in SmB6Phys. Rev. X 4, 031012 (2014).
  • Xu et al. [2016] Y. Xu, S. Cui, J. K. Dong, D. Zhao, T. Wu, X. H. Chen, K. Sun, H. Yao, and S. Y. Li, Bulk Fermi Surface of Charge-Neutral Excitations in SmB6 or Not: A Heat-Transport Study, Phys. Rev. Lett. 116, 246403 (2016).
  • Li et al. [2014] G. Li, Z. Xiang, F. Yu, T. Asaba, B. Lawson, P. Cai, C. Tinsman, A. Berkley, S. Wolgast, Y. S. Eo, D.-J. Kim, C. Kurdak, J. W. Allen, K. Sun, X. H. Chen, Y. Y. Wang, Z. Fisk, and L. Li, Two-dimensional Fermi surfaces in Kondo insulator SmB6Science 346, 1208 (2014).
  • Tan et al. [2015] B. S. Tan, Y.-T. Hsu, B. Zeng, M. C. Hatnean, N. Harrison, Z. Zhu, M. Hartstein, M. Kiourlappou, A. Srivastava, M. D. Johannes, T. P. Murphy, J.-H. Park, L. Balicas, G. G. Lonzarich, G. Balakrishnan, and S. E. Sebastian, Unconventional Fermi surface in an insulating state, Science 349, 287 (2015).
  • LaBarre et al. [2021] P. G. LaBarre, A. Rydh, J. Palmer-Fortune, J. A. Frothingham, S. T. Hannahs, A. P. Ramirez, and N. Fortune, Observation of Quantum Oscillations in The Low Temperature Specific Heat of SmB6preprint , arXiv:2111.03758 (2021).
  • Nomoto et al. [2022] T. Nomoto, S. Yamashita, H. Akutsu, Y. Nakazawa, and R. Kato, Systematic study on thermal conductivity of organic triangular lattice systems β\beta^{\prime}-X[Pd(dmit)2]2Phys. Rev. B 105, 245133 (2022).
  • Yamashita et al. [2022] M. Yamashita, Y. Sato, Y. Kasahara, S. Kasahara, T. Shibauchi, and Y. Matsuda, Resistivity and thermal conductivity of an organic insulator β\beta^{\prime}–EtMe3Sb[Pd(dmit)2]2Sci. Rep. 12, 9187 (2022).
  • Lewin et al. [2023] S. K. Lewin, C. E. Frank, S. Ran, J. Paglione, and N. P. Butch, A review of UTe2 at high magnetic fields, Rep. Prog. Phys. 86, 114501 (2023).
  • Suetsugu et al. [2024] S. Suetsugu, M. Shimomura, M. Kamimura, T. Asaba, H. Asaeda, Y. Kosuge, Y. Sekino, S. Ikemori, Y. Kasahara, Y. Kohsaka, M. Lee, Y. Yanase, H. Sakai, P. Opletal, Y. Tokiwa, Y. Haga, and Y. Matsuda, Fully gapped pairing state in spin-triplet superconductor UTe2Sci. Adv. 10, eadk3772 (2024).
  • Stankiewicz et al. [2019] J. Stankiewicz, M. Evangelisti, P. F. S. Rosa, P. Schlottmann, and Z. Fisk, Physical properties of SmxB6 single crystals, Phys. Rev. B 99, 045138 (2019).
  • Sato et al. [2019] Y. Sato, Z. Xiang, Y. Kasahara, T. Taniguchi, S. Kasahara, L. Chen, T. Asaba, C. Tinsman, H. Murayama, O. Tanaka, Y. Mizukami, T. Shibauchi, F. Iga, J. Singleton, L. Li, and Y. Matsuda, Unconventional thermal metallic state of charge-neutral fermions in an insulator, Nat. Phys. 15, 954 (2019).
  • Yamashita et al. [2011] S. Yamashita, T. Yamamoto, Y. Nakazawa, M. Tamura, and R. Kato, Gapless spin liquid of an organic triangular compound evidenced by thermodynamic measurements, Nat. Commun. 2, 275 (2011).
  • Bourgeois-Hope et al. [2019] P. Bourgeois-Hope, F. Laliberté, E. Lefrançois, G. Grissonnanche, S. R. de Cotret, R. Gordon, S. Kitou, H. Sawa, H. Cui, R. Kato, L. Taillefer, and N. Doiron-Leyraud, Thermal Conductivity of the Quantum Spin Liquid Candidate EtMe3Sb[Pd(dmit)2]2: No Evidence of Mobile Gapless Excitations, Phys. Rev. X 9, 041051 (2019).
  • Li et al. [2024] N. Li, M. T. Xie, Q. Huang, Z. W. Zhuo, Z. Zhang, E. S. Choi, Y. Y. Wang, H. Liang, Y. Sun, D. D. Wu, Q. J. Li, H. D. Zhou, G. Chen, X. Zhao, Q. M. Zhang, and X. F. Sun, Thermodynamics and heat transport in the quantum spin liquid candidates NaYbS2 and NaYbSe2Phys. Rev. B 110, 224414 (2024).
  • Ranjith et al. [2019] K. M. Ranjith, S. Luther, T. Reimann, B. Schmidt, P. Schlender, J. Sichelschmidt, H. Yasuoka, A. M. Strydom, Y. Skourski, J. Wosnitza, H. Kühne, T. Doert, and M. Baenitz, Anisotropic field-induced ordering in the triangular-lattice quantum spin liquid NaYbSe2Phys. Rev. B 100, 224417 (2019).
  • Dai et al. [2021] P.-L. Dai, G. Zhang, Y. Xie, C. Duan, Y. Gao, Z. Zhu, E. Feng, Z. Tao, C.-L. Huang, H. Cao, A. Podlesnyak, G. E. Granroth, M. S. Everett, J. C. Neuefeind, D. Voneshen, S. Wang, G. Tan, E. Morosan, X. Wang, H.-Q. Lin, L. Shu, G. Chen, Y. Guo, X. Lu, and P. Dai, Spinon Fermi Surface Spin Liquid in a Triangular Lattice Antiferromagnet NaYbSe2Phys. Rev. X 11, 021044 (2021).
  • Yu et al. [2017] Y. J. Yu, Y. Xu, L. P. He, M. Kratochvilova, Y. Y. Huang, J. M. Ni, L. Wang, S.-W. Cheong, J.-G. Park, and S. Y. Li, Heat transport study of the spin liquid candidate 1T-TaS2Phys. Rev. B 96, 081111 (2017).
  • Kratochvilova et al. [2017] M. Kratochvilova, A. D. Hillier, A. R. Wildes, L. Wang, S.-W. Cheong, and J.-G. Park, The low-temperature highly correlated quantum phase in the charge-density-wave 1T-TaS2 compound, npj Quantum Mater. 2, 42 (2017).
  • Imajo et al. [2019] S. Imajo, Y. Kohama, A. Miyake, C. Dong, M. Tokunaga, J. Flouquet, K. Kindo, and D. Aoki, Thermodynamic Investigation of Metamagnetism in Pulsed High Magnetic Fields on Heavy Fermion Superconductor UTe2J. Phys. Soc. Jpn. 88, 083705 (2019).
  • Ishihara et al. [2023] K. Ishihara, M. Roppongi, M. Kobayashi, K. Imamura, Y. Mizukami, H. Sakai, P. Opletal, Y. Tokiwa, Y. Haga, K. Hashimoto, and T. Shibauchi, Chiral superconductivity in UTe2 probed by anisotropic low-energy excitations, Nat. Commun. 14, 2966 (2023).
  • Bordelon et al. [2019] M. M. Bordelon, E. Kenney, C. Liu, T. Hogan, L. Posthuma, M. Kavand, Y. Lyu, M. Sherwin, N. P. Butch, C. Brown, M. J. Graf, L. Balents, and S. D. Wilson, Field-tunable quantum disordered ground state in the triangular-lattice antiferromagnet NaYbO2Nat. Phys. 15, 1058 (2019).
  • Vollmer [2009] M. Vollmer, Newton’s law of cooling revisited, Eur. J. Phys. 30, 1063 (2009).
  • Kakaç and Yener [1985] S. Kakaç and Y. Yener, Heat Conduction (Hemisphere Publishing Corporation, 1985).
  • Mills [1992] A. Mills, Heat Transfer (Irwin, 1992).
  • Chen et al. [2015] Q. Chen, R.-H. Fu, and Y.-C. Xu, Electrical circuit analogy for heat transfer analysis and optimization in heat exchanger networks, Appl. Energy 139, 81 (2015).
  • Bueno et al. [2012] B. Bueno, L. Norford, G. Pigeon, and R. Britter, A resistance-capacitance network model for the analysis of the interactions between the energy performance of buildings and the urban climate, Build. Environ. 54, 116 (2012).
  • Silva [2022] D. Silva, Modeling the Transient Response of Thermal Circuits, Appl. Sci. 1210.3390/app122412555 (2022).
  • [32] See the Supplementary Material for more details .

Supplementary material – Resolving the Thermal Paradox: Many-body localization or fractionalization?
Saikat Banerjee1,{}^{1,\,*}, Piers Coleman1,2,{}^{1,2,\,{\dagger}}

1 Center for Materials Theory, Rutgers University, Piscataway, New Jersey, 08854, USA
2 Hubbard Theory Consortium, Department of Physics,
Royal Holloway, University of London, Egham, Surrey TW20 0EX, UK
(Dated: )

I Impedance analysis

Here, we provide the derivation of time constants for the various thermal circuits relevant to the measurement of the specific heat and the thermal conductivity in the main text. The impedance for the first circuit in Fig. 1(a) in the main text is calculated as follows. First, we label the imaginary frequency by iωs-i\omega\rightarrow s. It is clear that the impedance for CfC_{\rm{f}}, and the thermal resistance RpfR_{\rm{pf}} is in series, which together is in parallel to the capacitance CpC_{\rm{p}}. This altogether yields a total impedance Z1(s)Z_{1}(s) as

Z1(s)=1sCp+1Rpf+1sCf=1sCp(s+1RpfCf)s+1RpfCp+1RpfCf,Z_{1}(s)=\frac{1}{sC_{\rm{p}}+\frac{1}{R_{\rm{pf}}+\frac{1}{sC_{\rm{f}}}}}=\frac{1}{sC_{\rm{p}}}\frac{\left(s+\frac{1}{R_{\rm{pf}}C_{\rm{f}}}\right)}{s+\frac{1}{R_{\rm{pf}}C_{\rm{p}}}+\frac{1}{R_{\rm{pf}}C_{\rm{f}}}}, (S1)

which gives a pole at

s0=1τ0=1RpfCp+1RpfCf.-s_{0}=\frac{1}{\tau_{0}}=\frac{1}{R_{\rm{pf}}C_{\rm{p}}}+\frac{1}{R_{\rm{pf}}C_{\rm{f}}}. (S2)

This is consistent with the time constant relevant to the specific heat measurement, as mentioned in Eq. (4) in the main text. Note that the other time constant is infinite [s=0s=0 in Eq. (S2)].

Impedance for thermal conductivity – Next, we compute the impedance for the second circuit in the main text [see Fig. 1(b)] in the same manner. In this case, though, two distinct time constants emerge. The impedance for the thermal conductivity measurements is calculated as

Z2(s)=1sCp+1ZR(s)+1Rp,where,ZR(s)=Rpf+1sCf+1Rf.Z_{2}(s)=\frac{1}{sC_{\rm{p}}+\frac{1}{Z_{R}(s)}+\frac{1}{R_{\rm{p}}}},\quad\text{where},\quad Z_{R}(s)=R_{\rm{pf}}+\frac{1}{sC_{\rm{f}}+\frac{1}{R_{\rm{f}}}}. (S3)

We now write the form of Z2(s)Z_{2}(s) explicitly to obtain the corresponding pole structure. The latter is obtained as

Z2(s)=Rp(CfRfRpfs+Rpf+Rf)CpCfRpRpfRfs2+s[Cf(Rp+Rpf)Rf+CpRp(Rpf+Rf)]+Rp+Rpf+Rf,Z_{2}(s)=\frac{R_{\rm{p}}(C_{\rm{f}}R_{\rm{f}}R_{\rm{pf}}s+R_{\rm{pf}}+R_{\rm{f}})}{C_{\rm{p}}C_{\rm{f}}R_{\rm{p}}R_{\rm{pf}}R_{\rm{f}}s^{2}+s[C_{\rm{f}}(R_{\rm{p}}+R_{\rm{pf}})R_{\rm{f}}+C_{\rm{p}}R_{\rm{p}}(R_{\rm{pf}}+R_{\rm{f}})]+R_{\rm{p}}+R_{\rm{pf}}+R_{\rm{f}}}, (S4)

which subsequently provides the two decay rates by solving for the roots of the denominator in Eq. (S4). Simplifying further, we obtain the roots as

s1\displaystyle s_{1} =(1Rp+1Rpf2Cp1Rpf+1Rf2Cf)2+1CpCfRpf21Rp+1Rpf2Cp1Rpf+1Rf2Cf,\displaystyle=\sqrt{\left(\frac{\frac{1}{R_{\rm{p}}}+\frac{1}{R_{\rm{pf}}}}{2C_{\rm{p}}}-\frac{\frac{1}{R_{\rm{pf}}}+\frac{1}{R_{\rm{f}}}}{2C_{\rm{f}}}\right)^{2}+\frac{1}{C_{\rm{p}}C_{\rm{f}}R_{\rm{pf}}^{2}}}-\frac{\frac{1}{R_{\rm{p}}}+\frac{1}{R_{\rm{pf}}}}{2C_{\rm{p}}}-\frac{\frac{1}{R_{\rm{pf}}}+\frac{1}{R_{\rm{f}}}}{2C_{\rm{f}}}, (S5a)
s2\displaystyle s_{2} =(1Rp+1Rpf2Cp1Rpf+1Rf2Cf)2+1CpCfRpf21Rp+1Rpf2Cp1Rpf+1Rf2Cf.\displaystyle=-\sqrt{\left(\frac{\frac{1}{R_{\rm{p}}}+\frac{1}{R_{\rm{pf}}}}{2C_{\rm{p}}}-\frac{\frac{1}{R_{\rm{pf}}}+\frac{1}{R_{\rm{f}}}}{2C_{\rm{f}}}\right)^{2}+\frac{1}{C_{\rm{p}}C_{\rm{f}}R_{\rm{pf}}^{2}}}-\frac{\frac{1}{R_{\rm{p}}}+\frac{1}{R_{\rm{pf}}}}{2C_{\rm{p}}}-\frac{\frac{1}{R_{\rm{pf}}}+\frac{1}{R_{\rm{f}}}}{2C_{\rm{f}}}. (S5b)

The denominator in Eq. (S4) can be written as s2+αs+βs^{2}+\alpha s+\beta, where we have

α=1Rpf(Rp+RpfRpCp+Rpf+RpfRpfCf),β=Rp+Rpf+RfCpCfRpRpfRf.\alpha=\frac{1}{R_{\rm{pf}}}\left(\frac{R_{\rm{p}}+R_{\rm{pf}}}{R_{\rm{p}}C_{\rm{p}}}+\frac{R_{\rm{pf}}+R_{\rm{pf}}}{R_{\rm{pf}}C_{\rm{f}}}\right),\quad\beta=\frac{R_{\rm{p}}+R_{\rm{pf}}+R_{\rm{f}}}{C_{\rm{p}}C_{\rm{f}}R_{\rm{p}}R_{\rm{pf}}R_{\rm{f}}}. (S6)

Slow and fast poles Now we analyze the structure of the poles at various limiting cases. Typically, the CfCpC_{\rm{f}}\gg C_{\rm{p}} in the various materials mentioned in the main text. Therefore, in this case, we have the β0\beta\to 0, and we have the fast pole at

s1=1τF=αRp+RpfRpRpfCpτFCp(RpRpfRp+Rpf).s_{1}=-\frac{1}{\tau_{\rm{F}}}=-\alpha\sim-\frac{R_{\rm{p}}+R_{\rm{pf}}}{R_{\rm{p}}R_{\rm{pf}}C_{\rm{p}}}\quad\Rightarrow\quad\tau_{\rm{F}}\sim C_{\rm{p}}\left(\frac{R_{\rm{p}}R_{\rm{pf}}}{R_{\rm{p}}+R_{\rm{pf}}}\right). (S7)

On the other hand, since s1s2=βs_{1}s_{2}=\beta, we have a slow pole leading to slow time-constant as

s2=1τSRp+Rpf+RfCpCf(RpRpfRf)(CpRpRpf(Rp+Rpf))=Rpf+Rp+RfCfRf(Rpf+Rp),τS=Cf(11Rf+1Rp+Rpf).s_{2}=-\frac{1}{\tau_{\rm{S}}}\sim\frac{R_{\rm{p}}+R_{\rm{pf}}+R_{\rm{f}}}{C_{\rm{p}}C_{\rm{f}}(R_{\rm{p}}R_{\rm{pf}}R_{\rm{f}})}\left(\frac{C_{\rm{p}}R_{\rm{p}}R_{\rm{pf}}}{-(R_{\rm{p}}+R_{\rm{pf}})}\right)=-\frac{R_{\rm{pf}}+R_{\rm{p}}+R_{\rm{f}}}{C_{\rm{f}}R_{\rm{f}}(R_{\rm{pf}}+R_{\rm{p}})},\quad\Rightarrow\quad\tau_{\rm{S}}=C_{\rm{f}}\left(\frac{1}{\frac{1}{R_{\rm{f}}}+\frac{1}{R_{\rm{p}}+R_{\rm{pf}}}}\right). (S8)

These two time constants are inherent to the thermal conductivity measurement circuits as explained in Eq. (5) in the main text. We now move on to deriving the time-dependent transient temperature profile in the sample using Laplace transformation.

Response in time domain We can formaly write the impedance in Eq. (S4) as following

Z2(s)=s+s0Cph(ss1)(ss2),wheres0=1Cf(1Rpf+1Rf),Z_{2}(s)=\frac{s+s_{0}}{C_{\rm{ph}}(s-s_{1})(s-s_{2})},\quad\text{where}\quad s_{0}=\frac{1}{C_{\rm{f}}}\left(\frac{1}{R_{\rm{pf}}}+\frac{1}{R_{\rm{f}}}\right), (S9)

and s1s_{1}, s2s_{2} are the poles as defined in Eqs. (S5a),(S5b). Simplifying further we rewrite Eq. (S9) as

Z2(s)=1Cp[s1+s0(ss1)(s1s2)s2+s0(ss2)(s1s2)].Z_{2}(s)=\frac{1}{C_{\rm{p}}}\left[\frac{s_{1}+s_{0}}{(s-s_{1})(s_{1}-s_{2})}-\frac{s_{2}+s_{0}}{(s-s_{2})(s_{1}-s_{2})}\right]. (S10)

Now, we utilize the electrical analogy of Laplace transformation to obtain the temperature profile as T(s)=Z2(s)I(s)T(s)=Z_{2}(s)I(s), where I(s)I(s) is the thermal current flowing in the sample. We adopt a crude approximation to the thermal current as

I(s)=I00est𝑑t=I0s,I(s)=I_{0}\int_{0}^{\infty}e^{-st}dt=\frac{I_{0}}{s}, (S11)

where I0I_{0} is the amplitude of the steady-state thermal current at a long time scale. Plugging this back into the temperature profile and simplifying it further we have

T(s)=I0Z2(s)s=I0Cp[s1+s0s1(s1s2)(1ss11s)s2+s0s2(s1s2)(1ss21s)].T(s)=I_{0}\frac{Z_{2}(s)}{s}=\frac{I_{0}}{C_{\rm{p}}}\left[\frac{s_{1}+s_{0}}{s_{1}(s_{1}-s_{2})}\left(\frac{1}{s-s_{1}}-\frac{1}{s}\right)-\frac{s_{2}+s_{0}}{s_{2}(s_{1}-s_{2})}\left(\frac{1}{s-s_{2}}-\frac{1}{s}\right)\right]. (S12)

Utilizing the Laplace transformation of eate^{at} as [eat]=1sa\mathcal{L}[e^{at}]=\tfrac{1}{s-a}, we perform the inverse Laplace transformation on Eq. (S12) to obtain the temperature profile as

T(t)=I0Cp[s1+s0s1(s1s2)(1es1t)+s2+s0s2(s1s2)(1es2t)].T(t)=\frac{I_{0}}{C_{\rm{p}}}\left[-\frac{s_{1}+s_{0}}{s_{1}(s_{1}-s_{2})}(1-e^{s_{1}t})+\frac{s_{2}+s_{0}}{s_{2}(s_{1}-s_{2})}(1-e^{s_{2}t})\right]. (S13)

where we labeled two constants A1A_{1}, and A2A_{2} as defined in Eq. (6) of the main text. They are defined in terms of the poles as

A1=s1+s0Cps1(s1s2),A2=s1+s0Cps2(s1s2),andτ1,2=1s1,2.A_{1}=-\frac{s_{1}+s_{0}}{C_{\rm{p}}s_{1}(s_{1}-s_{2})},\quad A_{2}=\frac{s_{1}+s_{0}}{C_{\rm{p}}s_{2}(s_{1}-s_{2})},\quad\text{and}\quad\tau_{1,2}=-\frac{1}{s_{1,2}}. (S14)