12
A Strange Metal from Gutzwiller correlations in infinite dimensions II: Transverse Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1 , Rok ˇ Zitko 2,3 , and B Sriram Shastry 1 1 Physics Department, University of California, Santa Cruz, California, 95060, 2 Joˇ zef Stefan Institute, Jamova 39, SI-1000 Ljubljana, Slovenia 3 Faculty for Mathematics and Physics, University of Ljubljana, Jadranska 19, SI-1000 Ljubljana, Slovenia (Dated: October 5, 2018) Using two approaches to strongly correlated systems, the extremely correlated Fermi liquid theory and the dynamical mean field theory, we compute the transverse transport coefficients, namely the Hall constants RH and Hall angles θH, and the longitudinal and transverse optical response of the U = Hubbard model in the limit of infinite dimensions. We focus on two successive low- temperature regimes, the Gutzwiller correlated Fermi liquid (GCFL) and the Gutzwiller correlated strange metal (GCSM). We find that the Hall angle cot θH is proportional to T 2 in the GCFL regime, while on warming into the GCSM regime it first passes through a downward bend and then continues as T 2 . Equivalently, RH is weakly temperature dependent in the GCFL regime, but becomes strongly temperature dependent in the GCSM regime. Drude peaks are found for both the longitudinal optical conductivity σxx(ω) and the optical Hall angles tan θH(ω) below certain characteristic energy scales. By comparing the relaxation rates extracted from fitting to the Drude formula, we find that in the GCFL regime there is a single relaxation rate controlling both longitudinal and transverse transport, while in the GCSM regime two different relaxation rates emerge. We trace the origin of this behavior to the dynamical particle-hole asymmetry of the Dyson self-energy, arguably a generic feature of doped Mott insulators. I. INTRODUCTION In a recent study 1 we have presented results for the longitudinal resistivity and low-temperature thermody- namics of the Hubbard model (with the repulsion pa- rameter U = ) in the infinite dimensional limit. In this limit, we can obtain the complete single-particle Green’s functions using two methods: the dynamic mean field theory (DMFT) 25 , and the extremely correlated Fermi liquid (ECFL) theory 6,7 , with some overlapping results and comparisons in Ref. [8]. These studies capture the non-perturbative local Gutzwiller correlation effects on the longitudinal resistivity ρ xx quantitatively 46 . A re- cent study by our group addresses the physically relevant case of two dimensions 9 , with important results for many variables discussed here. The present work extends the study of Ref. [1], using the ECFL scheme of Ref. [6], to the case of the Hall conductivity σ xy and the finite frequency (i.e. optical) conductivities. One goal is to further test ECFL with the exact DMFT results for these quantities which are more challenging to calculate. More importantly, however, by combining the various calculated conductivities we are able to uncover the emergence of two different transport relaxation times. In cuprate superconductors, various authors 1014 have commented on the different tempera- ture (T ) dependence of the transport properties in the normal phase. The cotangent Hall angles, defined as the ratio of the longitudinal conductivity σ xx and the Hall conductivity, cot(θ H )= σ xx xy , is close to quadratic as in conventional metals. Meanwhile, the longitudinal resistivity has unusual linear temperature dependence 15 . Understanding the ubiquitous T 2 behavior of cot(θ H ) in spite of the unconventional temperature dependence of the longitudinal resistivity is therefore quite important. In Ref. [1] we found that at the lowest tempera- tures the system is a Gutzwiller-correlated Fermi liquid (GCFL) with ρ xx T 2 . Upon warming one finds a regime with linear temperature dependence of the re- sistivity ρ xx 1 , which is reminiscent of the strange metal regime in the cuprate phase diagrams 15 . It is termed the Gutzwiller-correlated strange metal (GCSM) regime 1 . Previous studies 4,5 established the GCFL and GCSM regimes using the longitudinal resistivity. Here we fo- cus instead on the Hall constants R H = σ xy 2 xx and the Hall angles 5 , as well as on the optical conductivity 4 and optical Hall angles. In the GCFL regime, the pri- mary excitations are coherent quasiparticles that survive the Gutzwiller correlation, and there is a single transport relaxation time, as one would expect for a conventional Fermi liquid. Upon warming up into the GCSM regime, the longitudinal and transverse optical scattering rates become different. It appears that the existence of two separate scattering times is a generic characteristic of the GCSM regime. This work is organized as follows. First we summarize the Kubo formulas used to calculate the transport coeffi- cients in Sec. (II). We then revisit in Sec. (III) the famil- iar Boltzmann transport theory from which two separate relaxation times can be naturally derived. The results for the dc transport properties are presented in Sec. (IV) and those of optical conductivities in Sec. (V). In Sec. (VI) we interpret the two scattering times found in the GCSM regime through the particle-hole asymmetry of dynamical properties (spectral function) of the system. In conclusion we discuss the implication of this work for strongly correlated matter. arXiv:1705.01914v3 [cond-mat.str-el] 5 Jul 2017

Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

A Strange Metal from Gutzwiller correlations in infinite dimensions II: TransverseTransport, Optical Response and Rise of Two Relaxation Rates

Wenxin Ding1, Rok Zitko 2,3, and B Sriram Shastry1

1Physics Department, University of California, Santa Cruz, California, 95060,2 Jozef Stefan Institute, Jamova 39, SI-1000 Ljubljana, Slovenia

3Faculty for Mathematics and Physics, University of Ljubljana, Jadranska 19, SI-1000 Ljubljana, Slovenia(Dated: October 5, 2018)

Using two approaches to strongly correlated systems, the extremely correlated Fermi liquid theoryand the dynamical mean field theory, we compute the transverse transport coefficients, namely theHall constants RH and Hall angles θH , and the longitudinal and transverse optical response ofthe U = ∞ Hubbard model in the limit of infinite dimensions. We focus on two successive low-temperature regimes, the Gutzwiller correlated Fermi liquid (GCFL) and the Gutzwiller correlatedstrange metal (GCSM). We find that the Hall angle cot θH is proportional to T 2 in the GCFL regime,while on warming into the GCSM regime it first passes through a downward bend and then continuesas T 2. Equivalently, RH is weakly temperature dependent in the GCFL regime, but becomes stronglytemperature dependent in the GCSM regime. Drude peaks are found for both the longitudinaloptical conductivity σxx(ω) and the optical Hall angles tan θH(ω) below certain characteristic energyscales. By comparing the relaxation rates extracted from fitting to the Drude formula, we find thatin the GCFL regime there is a single relaxation rate controlling both longitudinal and transversetransport, while in the GCSM regime two different relaxation rates emerge. We trace the origin ofthis behavior to the dynamical particle-hole asymmetry of the Dyson self-energy, arguably a genericfeature of doped Mott insulators.

I. INTRODUCTION

In a recent study1 we have presented results for thelongitudinal resistivity and low-temperature thermody-namics of the Hubbard model (with the repulsion pa-rameter U =∞) in the infinite dimensional limit. In thislimit, we can obtain the complete single-particle Green’sfunctions using two methods: the dynamic mean fieldtheory (DMFT)2–5, and the extremely correlated Fermiliquid (ECFL) theory6,7, with some overlapping resultsand comparisons in Ref. [8]. These studies capture thenon-perturbative local Gutzwiller correlation effects onthe longitudinal resistivity ρxx quantitatively4–6. A re-cent study by our group addresses the physically relevantcase of two dimensions9, with important results for manyvariables discussed here.

The present work extends the study of Ref. [1], usingthe ECFL scheme of Ref. [6], to the case of the Hallconductivity σxy and the finite frequency (i.e. optical)conductivities. One goal is to further test ECFL with theexact DMFT results for these quantities which are morechallenging to calculate. More importantly, however, bycombining the various calculated conductivities we areable to uncover the emergence of two different transportrelaxation times. In cuprate superconductors, variousauthors10–14 have commented on the different tempera-ture (T ) dependence of the transport properties in thenormal phase. The cotangent Hall angles, defined as theratio of the longitudinal conductivity σxx and the Hallconductivity, cot(θH) = σxx/σxy, is close to quadraticas in conventional metals. Meanwhile, the longitudinalresistivity has unusual linear temperature dependence15.Understanding the ubiquitous T 2 behavior of cot(θH) inspite of the unconventional temperature dependence of

the longitudinal resistivity is therefore quite important.

In Ref. [1] we found that at the lowest tempera-tures the system is a Gutzwiller-correlated Fermi liquid(GCFL) with ρxx ∝ T 2. Upon warming one finds aregime with linear temperature dependence of the re-sistivity ρxx

1, which is reminiscent of the strange metalregime in the cuprate phase diagrams15. It is termedthe Gutzwiller-correlated strange metal (GCSM) regime1.Previous studies4,5 established the GCFL and GCSMregimes using the longitudinal resistivity. Here we fo-cus instead on the Hall constants RH = σxy/σ

2xx and

the Hall angles5, as well as on the optical conductivity4

and optical Hall angles. In the GCFL regime, the pri-mary excitations are coherent quasiparticles that survivethe Gutzwiller correlation, and there is a single transportrelaxation time, as one would expect for a conventionalFermi liquid. Upon warming up into the GCSM regime,the longitudinal and transverse optical scattering ratesbecome different. It appears that the existence of twoseparate scattering times is a generic characteristic of theGCSM regime.

This work is organized as follows. First we summarizethe Kubo formulas used to calculate the transport coeffi-cients in Sec. (II). We then revisit in Sec. (III) the famil-iar Boltzmann transport theory from which two separaterelaxation times can be naturally derived. The resultsfor the dc transport properties are presented in Sec. (IV)and those of optical conductivities in Sec. (V). In Sec.(VI) we interpret the two scattering times found in theGCSM regime through the particle-hole asymmetry ofdynamical properties (spectral function) of the system.In conclusion we discuss the implication of this work forstrongly correlated matter.

arX

iv:1

705.

0191

4v3

[co

nd-m

at.s

tr-e

l] 5

Jul

201

7

Page 2: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

2

II. KUBO FORMULAS

The transport properties of correlated materials canbe easily evaluated in the limit of infinite dimensionsbecause the vertex corrections are absent16. For dimen-sions d > 3, the longitudinal conductivity σxx is straight-forwardly generalized as the electric field remains a d-dimensional vector. The generalization is less clear forthe transverse conductivity and Hall constants, becausethe magnetic field is no longer a vector but rather arank-2 tensor defined through the electromagnetic ten-sor. Nevertheless, σxy can still be defined through suit-able current-current correlation functions.

The input to the transport calculation is the single-particle Green’s function G(ω,k), calculated in the fol-lowing within either ECFL or DMFT. The Kubo formu-las can be written as17,18

σxx = 2πq2e

∑k Φxxk

∫dω(−∂f(ω)

∂ω )ρ2G(ω,k), (1)

σxy/B =4π2q3

e

3

∑k Φxyk

∫dω(−∂f(ω)

∂ω )ρ3G(ω,k), (2)

where ρG(ω,k) = −ImG(ω,k)/π is the single-particlespectral function and qe = −|e| is the electron charge.Φxxk = (εxk)2 and Φxyk = (εyk)2εxxk − εykε

xkεxyk are called

transport functions, with εαk = ∂εk/∂kα and εαβk =∂2εk/∂kα∂kβ , εk being the energy dispersion. We set~ to 1.

It is more convenient to convert the multi-dimensionalk-sums into energy integrals:

σxx = σ02πD∫dεΦxx(ε)

Φxx(0)

∫dω(−∂f(ω)

∂ω )ρ2G(ω, ε), (3)

σxy/B = σ04π2Dqe

3

∫dεΦxy(ε)

Φxx(0)

∫dω(−∂f(ω)

∂ω )ρ3G(ω, ε),(4)

where Φxx(xy)(ε) =∑

k Φxx(xy)k δ(ε − εk), σ0 =

q2eΦxx(0)/D is the Ioffe-Regel-Mott conductivity, D is

half-bandwidth, and ρG(ω, ε) = ρG(ω,k) such that ε =εk. In d dimensions the transport functions on the Bethelattice are19

Φxx(ε) =1

3d(D2 − ε2)ρ0(ε), (5)

Φxy(ε) = − 1

3d(d− 1)ε(D2 − ε2)ρ0(ε), (6)

where ρ0(ε) = 2πD2

√D2 − ε2Θ(D − |ε|) is the non-

interacting density of states on the Bethe lattice and D isthe half bandwidth. Even though the transport functionresults indicate that σ vanishes as d → ∞, we can rede-fine the conductivities in this limit as the sum of all com-ponents: σL =

∑α σαα, σT =

∑α6=β Sgn[α− β]σαβ with

α(β) = 1, 2, . . . , d. More importantly, the d-dependencedirectly drops out when we compute the Hall constantRH = σxy/σ

2xx. For the rest of this work, we shall re-

define σxx and σxy via σL and σT considering that allcomponents of σL(T ) are equal so that both the d-factorand the constant factor drop out from the transport func-tions:

σxx = 3σL, Φxx(ε) = (D2 − ε2)ρ0(ε), (7)

σxy = 3σT , Φxy(ε) = −ε(D2 − ε2)ρ0(ε). (8)

III. TWO-RELAXATION-TIME BEHAVIOR INTHE BOLTZMANN THEORY

In Boltzmann theory, the transport properties canbe obtained by solving for the distribution functionin the presence of external fields from the Boltzmannequation20:

∂ δf

∂t− qe

~cv×B · ∂ δf

∂k+ v · qeE(t)

(−∂f

0

∂ε

)= L δf, (9)

where f is the full distribution function that needs tobe solved, f0 is the Fermi-Dirac distribution function,δf = f − f0, and Lδf represents the linearized collisionintegrals.

In the regime of linear response, we expand δfE,B inpowers of the external fields to second order as

δfE,B = δfE,0 +BδfE,1, (10)

where δfE,0 is the solution in the absence of magneticfields, and both δfE,0 and δfE,1 are linear in E. Inthe relaxation-time-approximation (RTA)21 we replace

the collision integrals as Lkδf → −δf/τ where τ is as-

sumed to be k-independent. However, LδfE,0 and LδfE,1

are in principle governed by different relaxation times, aspointed out by Anderson10,14. Writing

LδfE,0 → −δfE,0

τtr, LδfE,1 → −δf

E,1

τH, (11)

we obtain

σxx(ω) =ω2p

τtr1− iωτtr

, (12)

σxy(ω)/B =ω2pωc/B

4π2

τH1− iωτH

τtr1− iωτtr

, (13)

where

ω2p

4π=

∫ddk

(2π)d2q2ev

2x(−∂εf0), (14)

ωcB

= ω−2p

∫ddk

(2π)d2q3e(v2

x∂kyvy − vxvy∂kxvy)∂εf0, (15)

va = ∂kaε(k) is the velocity in direction a, ε(k) is energydispersion of the electrons and B = zB. Then the Hallangle is

tan θH(ω) =ωcπ

τH1− iωτH

. (16)

Therefore, the optical conductivities can be cast in theBoltzmann-RTA form as

σxx(0)

Re[σxx(ω)]= 1 + ω2τ2

tr, (17)

σxy(0)/B

Re[σxy(ω)]/B= 1 + ω2(τ2

tr + τ2H) + τ2

trτ2Hω

4, (18)

θH(0)

Re[θH(ω)]= 1 + ω2τ2

H . (19)

Page 3: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

3

n=0.7

n=0.75

n=0.8

n=0.85

0.000 0.005 0.010 0.015 0.0200.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

T/D

ρxx(σ0-1)

FIG. 1. Temperature dependence of the dc resistivity ρxx ofthe U = ∞ Hubbard model from DMFT (dashed lines) andECFL (solid symbols) for a range of electron densities n. Thehorizontal axis corresponds to absolute temperatures.

The dc and ac transport coefficients of a microscopictheory do not necessarily take the form of the BoltzmannRTA theory. In the rest of this work, we study both thedc and the real part of the ac transport coefficients, andconsider them as

<e[σxx(ω)] =σxx(0)

1 + τ2trω

2 +O(ω4), (20)

<e[tan θH(ω)/B] =tan θH(0)/B

1 + τ2Hω

2 +O(ω4). (21)

The relaxation times τtr and τH are extracted from thelow frequency part of <e[σxx(ω)] and <e[tan θH(ω)/B]by fitting to the above expressions. Although computingRe[θH(ω)] requires both real and imaginary parts of theoptical conductivities, we can make the approximationRe[θH(ω)] ' Re[σxy(ω)]/Re[σxx(ω)] when ω of concernis small. We expect τtr and τH to have similar tempera-ture and density dependence as σxx(0) and tan θH(0)/B.

IV. dc TRANSPORT

We now use the Kubo formulas to compute thetransport coefficients within the ECFL and DMFT ap-proaches. We plot the ECFL results as solid symbolsand the DMFT results as dashed lines using the samecolor for each density unless specified otherwise. As weshall demonstrate, the agreement between the DMFTand ECFL results follows the same qualitative trend forall quantities considered: it is better at lower tempera-tures, lower frequencies, and at lower density (higher holedoping).

We identify the GCFL and GCSM regimes, as well asthe cross-over scale TFL separating them, from the Tdependence of the longitudinal resistivity ρxx, shown inFig. (1). We identify the Fermi liquid temperature TFLusing the resistivity, rather than the more conventional

thermodynamic measures, such as heat capacity. Thelatter variables do actually give rather similar values, butthe resistivity seems most appropriate for this study. Ourdefinition is that up to and below TFL, the resistivityρxx ∼ T 2, while above TFL, ρxx displays a more complexset of T dependence as outlined in Ref. [1]. The Fermiliquid temperature has been quantitatively estimated inRef. [6]:

TFL ' 0.05×Dδα, (22)

where δ is the hole density δ = 1−n. The exponent α ∼1.39 within DMFT8; this is the value we will use below.α is somewhat greater for ECFL within the scheme usedin Ref. [6] and hence TFL given by DMFT is slightlyhigher than that by ECFL, as can also be seen in Fig. (1).Consequently as n increases, the the ECFL curves for ρxxlie above those from DMFT.

A. Hall constant

In Fig. (2), we show RH as a function of temperatureat different densities for low temperatures T < 0.02D(2a), and as a function of the hole density δ = (1 − n)at T = 0.002D, 0.005D, 0.01D (2b). The Hall constantis weakly temperature-dependent for T TFL, but itstarts to decrease on warming, as seen in Fig. (2a).

As a function of density δ the Hall constants from thetwo theories agree quite well, and are roughly linear withδ. The extrapolation to δ → 0 is uncertain from thepresent data. One might be tempted to speculate that itvanishes, since the lattice density of states is particle-holesymmetric. This question deserves further study withdifferent densities of states that break the particle-holesymmetry.

B. Cotangent of the Hall angle

The theoretical results for cotangent of the Hall angle(cot θH)B = (σxx/σxy)B are shown as a function of T 2

in Fig. (3a). We see that in DMFT as well as ECFL,the cot(θH) is linear in T 2 on both sides of a bend (orkink) temperature, which increases with increasing holedensity δ. However this kink is weaker in DMFT than inECFL. This bending was already noted in Fig. (5.a) ofRef. (9), within the 2-d ECFL theory. We may thus inferthat cot(θH) goes as QFLT

2 in the Fermi liquid regime,passes through a slight downward bend, and continues asQSMT

2 in the strange metal regimes, such that QFL >QSM . The difference, AFL−ASM , becomes smaller as δdecreases.

In order to characterize this kink more precisely, wedefine the downward bending regime by its onset temper-ature T−B , the crossing temperature of the two different

T 2 lines TB , and its ending temperature T+B . The tem-

peratures T−(+)B are determined by 5% deviation from

Page 4: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

4

n=0.7 n=0.75 n=0.8 n=0.85

0.005 0.010 0.015 0.0200.0

0.1

0.2

0.3

0.4

0.5

T/D

RH(qe/Vcell)

a)

T=0.002

T=0.005

T=0.01

0.15 0.2 0.25 0.3

0.1

0.2

0.3

0.4

0.5

δ

RH(qe/Vcell)

b)

FIG. 2. Temperature dependence of the Hall constants RH (2a) and RH at T = 0.002D, 0.005D, 0.01D as functions of the holedensity δ = 1 − n (2b) for both DMFT (dashed lines) and ECFL (solid symbols). RH is weakly T -dependent below TFL anddevelops stronger T -dependence in the GCSM regime. RH varies roughly linearly on δ at all three temperatures shown in (2b).

the T 2-fitting well below (above) TFL, and TB is welldefined as the crossing point of the two T 2-fittings. Weillustrate the kink and the determination of TB , T−B and

T+B at n = 0.7 for both ECFL in Fig. (3b) and DMFT in

Fig. (3c). In Fig. (3d), we show TB , T−B , T

+B and TFL

obtained from ECFL as functions of δ. We see that T−B is

identical to TFL, while TB and T+B are TFL plus some con-

stants with weak δ-dependence. We plot cot θHcot θH(T=TFL) as

functions of (T/TFL)2 for ECFL in Fig. (3e) and DMFTin Fig. (3f) to show the systematic evolution of the kinkswhen the density is varied.

C. Kink in cotangent of the Hall angle

There has been much interest in the quadratic T de-pendence of cot(θH) in the literature10,14. It is intrigu-ing that a kink in the plot of cot(θH) versus T 2 curves isseen in almost all experiments, although it appears to nothave been commented on earlier. Such bending is clearlyseen in experimental data Fig. (2) of Ref. [10], Fig. (4)of Ref. [22] and Fig. (3.c) of Ref. [23].

From Fig. (3.c) of Ref. [23] we estimate TB '100 K, 80 K, 70 K for LSCO at δ = 0.21, 0.17, 0.14respectively. These are comparable with the ECFL re-sults TB = 70 K, 60 K, 40 K at δ = 0.2, 0.175, 0.15,if we set D = 104 K. The trend of TB and the prefac-tor difference AFL−ASM also agrees with what we find,i.e., both TB and AFL − ASM decrease as δ is lowered.An increase of ASM at even higher temperatures is alsoobserved in Ref.[24], similar as what we find in Fig. (3a)above the GCSM regime.

It is notable that the bending temperatures TB in the-ory and in experiments are on a similar scale. It is there-fore of interest to explore this kink in cot(θH) more care-fully. From the perspective of the ECFL and DMFT the-

ories, we note that the kink represents one of the basiccrossovers discussed in Ref. [1], namely from the GCFLto GCSM regimes. It would be interesting to explorethis feature more closely in experiments, in particular tosee if the theoretically expected correlation between thecrossover in ρxx and cot(θH) finds support.

V. OPTICAL RESPONSE

A. Optical conductivity and the longitudinalscattering rates Γtr

In Fig. (4) we show the optical conductivity σxx(ω)as well as the quantity σxx(0)/σxx(ω) − 1, which betterpresents the approach to the zero frequency limit and isto be compared with the Boltzmann RTA form (Drudeformula) in Eq. (17). We display plots obtained fromboth ECFL (symbols) and DMFT (dashed lines) for fixedn = 0.8 and for three temperatures to show the genericbehavior at T < TFL, T ' TFL and T > TFL: T =0.002D (4a), T = 0.005D (4b) and T = 0.01D (4c).ECFL results agree well with the exact solution of DMFTwithin this temperature range.σxx(ω) shows a narrow Drude peak below TFL which

broadens as T increases and finally takes a form wellapproximated by a broad Lorentzian at T = 0.01 D.Correspondingly, (σxx(0)/σxx(ω)−1) is quadratic in fre-quency and can be fit to τ2

trω2 to extract the relaxation

time τtr. The ω2 regime has a width ∝ τ−1tr . The fitting

is performed at very small frequencies well within thisquadratic regime. At higher frequency, (σxx(0)/σxx(ω)−1) flattens out and creates a knee-like feature in-between.The flattening tendency decreases as T increases, and1/σxx(ω) grows monotonically. This knee-like featurethus becomes smoother as T increases and eventually islost for T > TFL. This trend is illustrated in Fig. (4d),

Page 5: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

5

n=0.7 n=0.75 n=0.8 n=0.825 n=0.85

0 1 2 3 4

0

1

2

3

4

T2/D2 (10-4)

cotθ

HμBB(D

-1)

a)

n=0.7, ECFL

TB

TB

+

TB

-

TFL

0 1 2 3 4

0.0

0.1

0.2

0.3

0.4

0.5

T2/D2(10-4)

cotθ

HμBB(D

-1)

b)

n=0.7, DMFT

TBTB

+

TB

-

TFL

0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0

0.1

0.2

0.3

0.4

T2/D2(10-4)

cotθ

HμBB(D

-1)

c)

×

×

××

×

TB × TFL

TB- TB

+

0.15 0.2 0.25 0.32

4

6

8

10

12

14

δ

T/D

(10-3)

d)

ECFL

n=0.7 n=0.75 n=0.8

n=0.825 n=0.85

0 1 2 3 40

1

2

3

4

(T/TFL)2

cotθ

H/cot

θH,TFL

e)

DMFT

n=0.7 n=0.75 n=0.8

n=0.825 n=0.85

0 1 2 3 40

1

2

3

4

(T/TFL)2

cotθ

H/cot

θH,TFL

f)

FIG. 3. Temperature dependence of the cotangent Hall angle cot θHB of both ECFL (symbols) and DMFT (dashed lines)shown as a function of T 2 (3a). The Hall angles cot θHB ∝ T 2 in the GCFL regime, passes through a slight downward bend(i.e., a kink), and continues as T 2 within the temperature range studied. The downward bending regime is characterized byits onset T−

B , the crossing of the two different T 2 lines TB , and its ending T+B . We illustrate the kink and the determination

of TB , T−B and T+

B at n = 0.7 for both ECFL (3b) and DMFT (3c). TB , T−B and T+

B obtained from the ECFL are shown as a

function of δ in (3d). We plot cot θHcot θH (T=TFL)

as functions of (T/TFL)2 for ECFL (3e) and DMFT (3f) to show the systematic

evolution of the kinks when the density varies.

where we normalize all curves of (σxx(0)/σxx(ω)− 1) bytheir corresponding τ2

tr, while the ω2 curve is shown as asolid blue line. All curves fall onto the ω2 line at smallfrequencies, and peal off at a frequency which increases

as T increases.

These scattering rates are shown as a function of tem-perature in Fig. (6a). The scattering rate Γ has a sim-ilar temperature dependence as the resistivity, i.e., a

Page 6: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

6

n=0.8,T=0.002D σECFL(ω)

σECFL(0)/σECFL(ω)-1

σDMFT(ω)

σDMFT(0)/σDMFT(ω)-1

0.00 0.02 0.04 0.06 0.08 0.100

100

200

300

400

ω/D

σ(ω

)(σ0)

a)

n=0.8,T=0.005D σECFL(ω)

σECFL(0)/σECFL(ω)-1

σDMFT(ω)

σDMFT(0)/σDMFT(ω)-1

0.00 0.02 0.04 0.06 0.08 0.10

0

10

20

30

40

50

60

ω/D

σ(ω

)(σ0)

b)

n=0.8,T=0.01D σECFL(ω)

σECFL(0)/σECFL(ω)-1

σDMFT(ω)

σDMFT(0)/σDMFT(ω)-1

0.00 0.02 0.04 0.06 0.08 0.100

5

10

15

20

25

ω/D

σ(ω

)(σ0)

c)

T=0.001

T=0.002

T=0.003

T=0.004

T=0.005

T=0.006

T=0.007

T=0.008

T=0.009

T=0.01

T=0.02

ω2

0.000 0.005 0.010 0.015 0.0200.0

0.5

1.0

1.5

2.0

2.5

3.0

ω/D

σ(0)/σ(ω

)-1

τ tr2

(10-4)

d)

FIG. 4. σxx(ω) and σ(0)/σ(ω)− 1 for n = 0.8 at T = 0.002D (4a), T = 0.005D (4b) and T = 0.01D (4c) for DMFT (dashedlines) and ECFL (solid symbols). The cyan solid lines are ω2 fitting near ω → 0. In (4d) we normalize σ(0)/σ(ω) − 1 curvescomputed from ECFL for various temperatures by τ2

tr with τtr obtained from the fits at small frequencies to the Drude formula.The solid blue line is a ω2 curve.

quadratic-T regime at low temperatures followed by alinear-T regime at higher temperatures.

B. Optical Hall angle and the transverse scatteringrates ΓH

In Fig. (5), we show the optical tangent Hall angletan θH(ω) and the quantity tan θH(0)/ tan θH(ω)−1. Wedisplay plots obtained from both ECFL (symbols) andDMFT (dashed lines) for fixed n = 0.8 and for threetemperatures to show the generic behavior at T < TFL,T ' TFL and T > TFL: T = 0.002D (5a), T = 0.005D(5b) and T = 0.01D (5c). The ECFL results agree wellwith those from DMFT within this temperature range.

Just like σxx(ω), tan θH(ω) possesses a narrow Drudepeak below TFL that broadens in a similar way withincreasing temperature. (tan θH(0)/ tan θH(ω) − 1) isquadratic in frequency and we fit τ2

Hω2 to extract the

transverse relaxation time τH . The ω2 regime, however,has a very narrow, weakly T -dependent width which is

about 0.003 D. The relaxation time τH is extracted byfitting within this very low frequency range. Above thisenergy a flattening behavior, similar to that in the opti-cal conductivity, takes place at low temperatures. Athigher temperatures and lower hole density, a power-law behavior with an exponent that increases with Tgradually replaces the flattening out behavior. Such atendency is visible in Figs. (5d) and (5e), where all(tan θH(0)/ tan θH(ω)−1) curves are normalized by theircorresponding τ2

H .

In Fig. (6b) we show ΓH (defined as ΓH ≡ τ−1H ) for var-

ious densities and temperatures obtained from the Drudeformula fitting. Their T -dependence is quadratic for bothGCFL and GCSM regimes.

C. Emergence of two relaxation times

In Fig. (6c), we show ΓH/Γtr as a function of tempera-ture. At all densities considered this ratio behaves differ-ently for T below and above TFL. Below TFL, the ratio

Page 7: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

7

n=0.8, T=0.002D

tanθH(ω)(μBB)-1 (D-1)

tanθHDMFT(ω)

0.000 0.005 0.010 0.015 0.020

0

5

10

15

20

25

30

ω/D

tanθ

H(ω

)(μBB)-1(D

-1)

tanθH (0)

tanθH (ω)-1

tanθHDMFT (0)

tanθHDMFT (ω)

-1

0

100

200

tanθ

H(0)/tanθ

H(ω

)-1

a)

n=0.8, T=0.005D

tanθH(ω) tanθHDMFT(ω)

0.000 0.005 0.010 0.015 0.020

0

1

2

3

4

5

6

ω/D

tanθ

H(ω

)

tanθH (0)

tanθH (ω)-1 tanθH

DMFT (0)

tanθHDMFT (ω)

-1

0

10

20

30

tanθ

H(0)

tanθ

H(ω

)-1

b)

n=0.8, T=0.01D tanθH(ω) tanθHDMFT(ω)

0.000 0.005 0.010 0.015 0.020

0.0

0.5

1.0

1.5

2.0

ω/D

tanθ

H(ω

)

tanθH (0)tanθH (ω)

-1tanθH

DMFT (0)

tanθHDMFT (ω)

-1

0

2

4

6

tanθ

H(0)/tanθ

H(ω

)-1

c)

n=0.7 T=0.001

T=0.004

T=0.007

T=0.01

T=0.013

T=0.016

T=0.019

0.000 0.005 0.010 0.015 0.0200.0

0.2

0.4

0.6

0.8

1.0

ω/D

(tan

θH(0)/tanθ

H(ω

)-1)/τH2(10-4)

d)

n=0.8 T=0.001

T=0.002

T=0.003

T=0.004

T=0.005

T=0.006

0.000 0.005 0.010 0.015 0.0200.0

0.2

0.4

0.6

0.8

1.0

ω/D

(tan

θH(0)/tanθ

H(ω

)-1)/τH2(10-4)

e)

FIG. 5. Optical Hall angles tan θH(ω) (blue) and tan θH(0)/ tan θH(ω)−1(red) shown for n = 0.8, T = 0.002D (5a), T = 0.005D(5b) and T = 0.01D (5c) for DMFT (dashed lines) and ECFL (solid symbols). The cyan solid lines are ω2 fitting near ω → 0.[tan θH(0)/ tan θH(ω) − 1]/τ2

H obtained from ECFL shown for n = 0.7 (5d) and n = 0.8 (5e). Drude peaks are found to benarrow (note the different horizontal axis scale compared to Fig. 3).

ΓH/Γtr ' 0.5 remains essentially constant, and hencethe optical transport is dominated by a single scatter-ing rate. Once TFL is crossed, however, ΓH/Γtr becomesstrongly T -dependent. This indicates that there are tworelaxation times in the GCSM regime. This is possi-ble since the quasiparticles are no longer well definedfor T > TFL, and different frequency regimes present

in the spectral functions contribute differently to the tworelaxation times. In Fig. (6d), we plot ΓH/Γtr versusthe rescaled temperature T/TFL to illustrate the clearlydistinct behavior below and above TFL.

Page 8: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

8

n=0.7

n=0.75

n=0.8

n=0.85

0.000 0.005 0.010 0.015 0.0200.00

0.01

0.02

0.03

0.04

0.05

T/D

Γtr(D

)

a)

n=0.7

n=0.75

n=0.8

n=0.85

0.000 0.005 0.010 0.015 0.020

0.00

0.01

0.02

0.03

0.04

T/D

ΓH(D

)

b)

n=0.7

n=0.75

n=0.8

n=0.85

0.000 0.005 0.010 0.015 0.0200.4

0.6

0.8

1.0

1.2

1.4

T/D

ΓH/Γtr

c)

n=0.7

n=0.75

n=0.8

n=0.85

0.0 0.5 1.0 1.5 2.0 2.50.45

0.50

0.55

0.60

0.65

0.70

0.75

0.80

T/TFL

ΓH/Γtr

d)

FIG. 6. Longitudinal relaxation rate Γtr extracted by fitting σxx(ω) by the Drude formula (6a), transverse relaxation time ΓHextracted from θH(ω) (6b), their ratio ΓH/Γtr as functions of T (6c) and as functions of scaled temperature T/TFL (6d). Allthe relaxation rates are extracted from the ECFL optical response results.

VI. ANALYSIS

We begin by analyzing the exact formulas for the con-ductivities σxx, σxy of Eqs. (3) and (4), following Ref. [25]and [6] within ECFL theory where more analytic insightis available.

It has long been noted that the particle-hole asymme-try of the spectral function is one of the characteristic fea-tures of strongly correlated systems26,29–34. The dynamicparticle-hole transformation is defined by simultaneouslyinverting the wave vector and energy in ρG(k, ω) relative

to the chemical potential µ as (k, ω)→ −(k, ω), with k =

k−kF 26. In the limit of d→∞, we ignore the k part ofthe transformation. Consequently, the dynamic particle-hole asymmetry solely stems from the asymmetry of theself-energy spectral function ρΣ(ω, T ) = −ImΣ(ω, T )/π.Instead of analyzing ρG, we can simply focus on ρΣ since

ρG =ρΣ

(ω + µ− ε−ReΣ)2 + π2ρ2Σ

. (23)

=1

π

B(ω, T )

(A(ω, T )− ε)2 +B2(ω, T )(24)

where

A(ω, T ) = ω + µ−ReΣ(ω, T ), (25)

B(ω, T ) = πρΣ(ω, T ) = −ImΣ(ω, T ). (26)

Then we approximate the exact equations (3) and (4)by their asymptotic values at low enough T, followingRef. [6]. The idea is to first integrate over the bandenergy ε viewing one of the powers of ρG as a δ functionconstraining ε→ A(ω, T ). This gives

σxx = σ0DΦxx[0]

∫dω(−f ′)Φxx[A(ω,T )]

B(ω,T ) , (27)

σxy = σ0DqeΦxx[0]

∫dω(f ′)

(∂2ωΦxy[A(ω,T )]

3 + Φxy [A(ω,T )]2(B(ω,T ))2

),

(28)

The first term in Eq. (28) turns out to be negligible com-pared to the second, and hence we will ignore it. Next,we track down the electronic properties that give rise toa second relaxation time using the above asymptotic ex-pressions.

To the lowest order of approximation at low temper-atures, we can make the substitution f ′(ω) → −δ(ω) in

Page 9: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

9

cot θH,0 n=0.7 n=0.75 n=0.8 n=0.85

0 1 2 3 40

1

2

3

4

5

T2/D2(10-4)

cotθ

H

0 0.01 0.02

0

2

4

T/D

FIG. 7. Zeroth order asymptotic cotangent Hall anglescot θH,0 plotted as functions of T 2 (main panel, symbols) com-pared with the exact results (dashed lines) and as functionsof T (inset).

Eq. (27) and (28), which gives

cot θH,0/B =2B(0, T )

qeA(0, T ). (29)

We show cot θH,0 in Fig. (7). When plotted as a func-tion of T 2 as shown in the main panel of Fig. (7), cot θH,0(solid symbols) is in good agreement with the exact re-sults (dashed lines) both qualitatively, i.e., showing akink-like feature, and quantitatively except for relativelyhigh temperatures and densities. However, when it isplotted as a function of T (inset of Fig. (7)), we findthat the ”kink” is actually the crossover from a T 2 be-havior to a linear-T behavior and cot θH,0 follows theT -dependence of ρxx. The lowest order approximation isinsufficient to capture and to understand the second T 2

regime. Therefore, we pursue more accurate asymptoticexpressions of Eqs. (27) and (28). Following Ref. [5]and [1], we do the following small frequency expansion:

Φxx(xy)[A(ω, T )] = Φxx(xy)[A0]

+ Φxx(xy)′[A0]A1 ω + . . . ,(30)

B(ω, T ) = B0 +B1ω +B2ω2 + . . . , (31)

where A0 and A1 is given by the expansion

A(ω, T ) = A0 +A1ω + . . . , (32)

Recall that A1 = Z−1, it is therefore large. In order toprovide further context to these coefficients Bn and toconnect with earlier discussions of the self energy, it isuseful to recall a useful and suggestive expression for theimaginary self energy exhibiting particle-hole asymmetryat kF at low ω (e.g. see Eq. (28) in Ref. [8])

−=mΣ(ω, T ) ∼ π (ω2 + π2T 2)

ΩΣ(T )

(1− ω

), (33)

where ΩΣ behaves as ∼ Z2 in the low-T Fermi liquidregime. The scale ∆ breaks the particle-hole symmetryof the leading term.

The variation of ΩΣ and ∆ in the GCSM regime isillustrated below in Fig. (8). Expanding this expression

at low ω we identify the coefficients B0 = π π2T 2

ΩΣ(T ) , B1 =

−B0

∆ , B2 = πΩΣ

, all of which are numerically verified tobe valid for all temperatures we study in this work. Thenegative sign of B1 is easily understood.

Now we keep B(ω, T ) to O(ω2) and A(ω, T ) to O(ω),which are the lowest orders required to capture all im-portant features of the exact results. Then Eq. (27) and(28) can be simplified as

σxx 'σ0F

01

D2B0(D2 −A2

0)3/2(

1− 3π2F 22

F 01

T 2A0A1

∆(D2 −A20)

),

(34)

σxy/B 'σ0qeF

02

2D2B20

A0(D2 −A20)3/2

×(

1 +π2F 2

3

F 02

T 2A1

∆A0(1− 3A2

0

D2 −A20

)).

(35)

The coefficients are defined as27

Fnm =π

4

∫ ∞−∞

dx

cosh2(πx/2)

xn

(1 + x2)m. (36)

Using Eqs. (34), (35) and [27], we can write

σxx ' σxx,0(1− αxx), (37)

σxy ' σxy,0(1− αxy), (38)

with

σxx,0 = σ0(D2 −A2

0)3/2

D2

0.822467

B0, (39)

σxy,0/B = σ0qeA0(D2 −A2

0)3/2

D2

0.355874

B20

, (40)

αxx =A1A0

D2 −A20

3.98598× T 2

∆, (41)

αxy = −A1

( 1

A0− 3A0

D2 −A20

)2.12075× T 2

∆. (42)

σxx,0 agrees with previous works1,6. αxx(xy) are relativecorrections due to ∆ and A1 comparing to σxx(xy),0. Nu-merical results of αxx and αxy are shown in Fig. (9a).We find that |αxx| is less than 5% even at the highesttemperature. However, αxy becomes O(1) in the GCSMregime. Therefore, we obtain the following asymptotictan θH by omitting αxx:

cot(θH) ' cot θH,0(1− αxy)

, (43)

cot θH,0/B = qeB0

0.432691A0. (44)

Page 10: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

10

n=0.7 n=0.75 n=0.8 n=0.85

0.000 0.005 0.010 0.015 0.0200.00

0.05

0.10

0.15

0.20

0.25

T/D

ΩΣ

a)

n=0.7 n=0.75 n=0.8 n=0.85

0.000 0.005 0.010 0.015 0.0200.00

0.05

0.10

0.15

0.20

0.25

T/D

Δ

b)

FIG. 8. Coefficients of the small frequency expansion of the ECFL Dyson self-energy ΩΣ (8a) and ∆ (8b) plotted as functionsof temperature.

We show ρxx and cot(θH) computed from the asymp-totic expressions Eq. (37) and (38) in Fig. (9). Theasymptotic values are denoted by crosses whereas the re-sults of Eq. (3) and (4) are denoted by solid circles.The numerical results of Eq. (43) recover the second T 2

regime.Therefore, we find that the αxy term due to the higher

order terms of A(ω, T ) and B(ω, T ) gives rise to the sec-ond T 2 regime of cot(θH). Typically such correction issmall, such as is the case of αxx. The significant differ-ence between αxx and αxy is understood by examiningEq. (41) and (42) more closely. Both αxx and αxy are∝ A1T

2/∆ with slightly different constant factors. SinceA0 D and almost independent of T , we can ignorethe 3A0(D2−A2

0)−1 term of αxy. Hence the difference ismostly determined by a factor

αxy/αxx ∼ A−20 , (45)

which greatly enhances αxy.In the GCFL regime, αxy is negligible, the coefficient

of the T 2 behavior is

QFL =B cot(θH)

T 2' π3

0.432691× qeA0ΩΣ(T → 0). (46)

ΩΣ(T ) is almost a constant in the GCFL regime hence ap-proximated by its zero temperature value ΩΣ(T → 0)28.In the GCSM regime, both ΩΣ and ∆ becomes linear-in-T :

ΩΣ(T ) ' Ω0(T + TΩ), (47)

∆(T ) ' ∆0(T + T∆), (48)

where Ω0(∆0) and TΩ(∆) are fitting parameters28. Bykeeping only the constant term we obtain

QSM 'π3

0.432691× qeA0Ω0(T∆ + TΩ). (49)

We compare the actual QFL and QSM with Eq. (46) and(49) in Fig. (10).

According to the above analysis, the second T 2 behav-ior of cot(θH) is due to the combination of two things:

• the dynamic particle-hole anti-symmetric compo-nent of ρΣ(ω) characterized by the energy scale∆. Its contribution to transport becomes impor-tant when πT becomes comparable to ∆;

• the particular form of the transverse transportfunction Φxy(ε) that causes Φxy′[A0]/Φxy[A0] ∝A−1

0 . Without this factor, αxy would be negligibleas αxx. This particular form of Φxy(ε) is due to theparticle-hole symmetry of the bare band structure.

VII. DISCUSSION

We have shown that Hall constants, Hall angles, opti-cal conductivities, and optical Hall angles calculated byECFL agree reasonably well with the DMFT results. Thedifferences tend to increase at higher densities and highertemperatures as noted earlier6.

We focused on the differences in the behavior aboveand below the Fermi liquid temperature scale TFL, i.e.,from the GCFL regime to the GCSM regime. BelowTFL, both ρxx and cot(θH) ∝ T 2. Equivalently, RHhas very weak T -dependence since RH = ρxx/ cot(θH).When T > TFL, however, cot(θH) passes through a slightdownward bend and continues as T 2 whereas ρxx ∝ T .The significance of the downward bend is that it signalsthe crossover to the strange metal regime from the Fermiliquid regime.

We explored the long-standing two-scattering-rateproblem by calculating both the optical conductivitiesand optical Hall angles, and the corresponding scatter-ing rates. Below TFL, both σxx(ω) and tan θH(ω) ex-hibit Drude peaks, which is a manifestation of transportdominated by quasiparticles. The corresponding scatter-ing rates can be extracted by fitting to the Drude for-mula in the appropriate frequency range. Above TFL,the Drude peak for σxx(ω) becomes broadened, i.e.,σxx(0)/σxx(ω) − 1 ∼ ω2 for an even larger range thatkeeps growing with increasing temperature. In this case,

Page 11: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

11

αxy n=0.7 n=0.75 n=0.8 n=0.85

αxx

0.000 0.005 0.010 0.015 0.020

0.0

0.2

0.4

0.6

0.8

T/D

αxy,αxx

a)

××××××××××××××

××××××

×××××××××××

××××××××

×

×××××××

×××××

×××××

×××

××××

××××

××××

××××

××××

Eq. (35)

× n=0.7

× n=0.75

× n=0.8

× n=0.85

0.000 0.005 0.010 0.015 0.020

0.00

0.05

0.10

0.15

0.20

0.25

0.30

T/D

ρxx

(σ0-1)

b)

Eq. (35)(36) × n=0.7 × n=0.75 × n=0.8 × n=0.85

×××××××××× × × × × × × × × × ××××××××

××× × × × × × × × × × ×

××××××××

××× × × × × × ×

××

×

×××××××

×

×

×

0 1 2 3 4

0

1

2

3

4

T2/D2 (10-4)

cotθ

HμBB(D

-1)

c)

FIG. 9. αxx (dashed lines) and αxy (solid symbols) (9a), ρxx (9b), cot(θH) (9c) computed from Eq. (37) and (38) using ECFLresults. The asymptotic values are denoted by crosses whereas the ECFL results of Eq. (3) and (4) are denoted by solid circles.

×××

×

×××

×

QFL

QSM

× Eq.(46)

× Eq.(48)

0.15 0.20 0.25 0.30

0

1

2

3

4

δ

QFL(SM)(104)

FIG. 10. Eq. (46) and (49) (crosses) compared with QFL andQSM (solid circles) obtained by fitting the exact cot(θH).

fitting to the Drude formula is still valid, and the scatter-ing rate shows similar trends as a function of temperatureas the dc resistivity. For θH(ω), the Drude peak rangeis very narrow, but nonetheless persists for all tempera-tures that we study in this work. Similarly, the extractedscattering rate ΓH shows similar trends as a function oftemperature as the dc Hall angle. At lower dopings andhigher temperatures, it seems possible that the Drudepeaks of θH(ω) would disappear and the fractional powerlaw would stretch down to nearly ω = 0.

By comparing the two optical scattering rates through

their ratio, ΓH/Γtr, we clearly demonstrated that ΓHand Γtr are equivalent below TFL, but that they quicklybecome two distinguishable quantities when the systemcrosses over into the strange-metal region.

By carefully examining the asymptotic expressions ofσxx and σxy we established that the different temper-ature dependence of cot(θH) in the GCSM regime isgoverned by a correction caused by both the dynamicalparticle-hole anti-symmetric component of ρΣ(ω) and theparticle-hole symmetry of the bare band structure. Thiscorrection is turned on when T becomes comparable to∆, the characteristic energy scale of the anti-symmetriccomponents of ρΣ(ω).

It would be useful to examine the bend in cot(θH) moreclosely in experiments in cuprate materials, where such afeature is apparently widely prevalent but seems to haveescaped comment so far. In particular, one would like tounderstand better if the longitudinal resistivity and thecotangent Hall angle show simultaneous signatures of acrossover, as the theory predicts.

VIII. ACKNOWLEDGEMENTS

The work at UCSC was supported by the U.S. Depart-ment of Energy (DOE), Office of Science, Basic EnergySciences (BES) under Award # DE-FG02-06ER46319.RZ acknowledges the financial support from the Slove-nian Research Agency (research core funding No. P1-0044 and project No. J1-7259).

1 W. Ding, R. Zitko, P. Mai, E. Perepelitsky, and B. S.Shastry, arXiv:1703.02206v2.

2 M. Walter and D. Vollhardt, Phys. Rev. Lett. 62, 324(1989).

3 A. Georges, G. Kotliar, W. Krauth and M. J. RozenbergRev. Mod. Phys. 68, 13 (1996).

4 X. Deng, J. Mravlje, R. Zitko, M. Ferrero, G. Kotliar,and A. Georges, Phys. Rev. Lett. 110, 086401 (2012),arXiv:1210.1769.

5 W. Xu, K. Haule, and G. Kotliar, Phys. Rev. Lett. 111,036401 (2013), arXiv:1304.7486.

Page 12: Transport, Optical Response and Rise of Two Relaxation Rates · Transport, Optical Response and Rise of Two Relaxation Rates Wenxin Ding 1, Rok Zitko 2;3, and B Sriram Shastry 1Physics

12

6 B. S. Shastry and E. Perepelitsky, Phys. Rev. B 94, 045138(2016), arXiv:1605.08213.

7 B. S. Shastry, Phys. Rev. Lett. 107, 056403 (2011),arXiv:1102.2858.

8 R. Zitko, D. Hansen, E. Perepelitsky, J. Mravlje,A. Georges, and B. S. Shastry, Phys. Rev. B 88, 235132(2013), arXiv:1309.5284.

9 B. S. Shastry and P. Mai, arXiv:1703.08142 (2017).10 T. Chien, Z. Wang, and N. Ong, Phys. Rev. Lett. 67, 2088

(1991).11 Y. Ando, Y. Kurita, S. Komiya, S. Ono and K. Segawa,

Phys. Rev. Letts. 92, 197001 (2004).12 F. F. Balakirev, J. B. Betts, A. Migliori, I. Tsukada, Y.

Ando, and G. S. Boebinger, Phys. Rev. Letts. 102, 017004(2009).

13 J. Takeda,T. Nishikawa, M. Sato, Physica C 231, 293(1994). See esp. Fig. (4).

14 P. W. Anderson, Phys. Rev. Lett. 67, 2092 (1991).15 Y. Ando, S. Komiya, K. Segawa, S. Ono, and Y. Kurita,

Phys. Rev. Lett. 93, 267001 (2004), arXiv:0403032 [cond-mat].

16 A. Khurana, Phys. Rev. Lett. 64, 1990 (1990).17 T. Pruschke, D. L. Cox, and M. Jarrell, Phys. Rev. B 47,

3553 (1993).18 P. Voruganti, A. Golubentsev, and S. John, Phys. Rev. B

45, 13945 (1992).19 L.-F. Arsenault and A.-M. S. Tremblay, Phys. Rev. B 88,

205109 (2013), arXiv:1305.6999.20 J. M. Ziman, “Electrons and phonons: the theory of trans-

port phenomena in solids,” (1960).21 D. Feng and G. Jin, in Introd. to Condens. Matter Phys.

(WORLD SCIENTIFIC, 2005) pp. 199–229.22 H. Y. Hwang, B. Batlogg, H. Takagi, H. L. Kao, J. Kwo,

R. J. Cava, J. J. Krajewski, and W. F. Peck, Jr., Phys.Rev. Lett. 72, 2636 (1994).

23 Y. Ando, Y. Kurita, S. Komiya, S. Ono, and K. Segawa,Phys. Rev. Lett. 92, 197001 (2004).

24 S. Ono, S. Komiya, and Y. Ando Phys. Rev. B 75, 024515(2007).

25 P. Voruganti, A. Golubentsev and S. John, Phys. Rev. B45, 13945 (1992).

26 B. S. Shastry, Phys. Rev. Lett. 109, 067004 (2012).27 Here we give numerical values of Fnm here, F 0

1 = π2/12 =0.822467, F 2

2 = π2/24−ζ(3)/4 = 0.110719, F 02 = π2/24+

ζ(3)/4 = 0.711748, F 23 = π2/96 − π4/960 + ζ(3)/16 =

0.0764691, where ζ(3) = 1.20206 . . . is the Reimann zetafunction.

28 Here we give numerical values of ΩΣ(T → 0), Ω0, ∆, TΩ

and T∆ for n = 0.7, 0.75, 0.8, 0.85 respectively:n 0.7 0.75 0.8 0.85ΩΣ(T → 0) 0.194326 0.10234 0.0443113 0.0135004Ω0 5.99932 4.97066 3.79492 2.52377∆ 5.15313 5.58921 5.97819 6.05317TΩ 0.0257944 0.0143857 0.00607066 0.000456418T∆ 0.0346638 0.0248982 0.0171223 0.0118911

29 Ch. Renner, B. Revaz, J.-Y. Genoud, K. Kadowaki, Ø.Fischer, Phys. Rev. Lett. 80, 149 (1998).

30 P. W. Anderson, N. P. Ong, J. Phys. Chem. Solid 67, 1(2006).

31 T. Hanaguri, C. Luplen, Y. Kohsaka, D.-H. Lee, M.Azuma, M. Takano, H. Takagi, J. C. Davis, Nature 430,1001 (2004).

32 A. N. Pasupathy, A. Pushp , K. K. Gomes, C. V. Parker, J.Wen, Z. Xu, G. Gu, S. Ono, Y. Ando, A. Yazdani, Science320, 196 (2008).

33 P. A. Casey, J. D. Koralek, N. C. Plumb, D. S. Dessau, P.W. Anderson, Nat. Phys. 4, 210 (2008).

34 G.-H. Gweon, B. S. Shastry, G. D. Gu, Phys. Rev. Lett.107, 056404 (2011).