License: CC BY 4.0
arXiv:2402.10878v2 [cond-mat.str-el] 22 Mar 2024

A model of non-Fermi liquid with power law resistivity: strange metal with a not-so-strange origin.

Patrick A. Lee Department of Physics, Massachusetts Institute of Technology, Cambridge, MA, USA
Abstract

We construct a model which exhibits resistivity going as a power law in temperature T𝑇Titalic_T, as Tαsuperscript𝑇𝛼T^{\alpha}italic_T start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT down to the lowest temperature. There is no residual resistivity because we assume the absence of disorder and momentum relaxation is due to umklapp scattering. Our model consists of a quantum spin liquid state with spinon Fermi surface and a hole Fermi surface made out of doped holes. The key ingredient is a set of singular 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT modes living on a ring in momentum space. Depending on parameters, α𝛼\alphaitalic_α may be unity (strange metal) or even smaller. The model may be applicable to a doped organic compound, which has been found to exhibit linear T resistivity. We conclude that it is possible to obtain strange metal behavior starting with a model which is not so strange.

I Introduction

The notion of strange metals emerged out of transport measurements from the early days of cuprate superconductivity and has been applied broadly over the years to instances where linear in temperature resistivity is observed, whether it is low temperature, high temperature and with or without a finite residual resistivity due to disorder. In this broad sense, strange metal behavior has been seen in many materials and has attracted a great deal of attention from the community, as summarized nicely in recent reviews  [1, 2]. The strange metal is usually associated with the violation of Landau’s theory of Fermi liquid, and believed to be driven by strong correlation physics. In this paper we will restrict the use of the term strange metal to low temperatures and in the absence of disorder. The high temperature linear resistivity anomaly is associated with violations of the Mott-Ioffe-Regel limit and clearly requires a separate set of physical input. [2] On the opposite end, while there are examples where linear T resistivity survives to low temperatures, in many cases the linear regime does not extend above the residual resistivity beyond a value comparable to the residual resistivity itself. Examples of this behavior include the overdoped [3] and electron doped cuprates [4]. While this phenomenon is not understood and is of great interest, the effect of disorder is likely to be strongly relevant and a different set of explanations may be required. Furthermore, as a matter of principle, Landau’s Fermi liquid theory refers to the clean case. In this paper we will not discuss models with disorder [5, 6, 7, 8]. We will focus on the situation where a power law resistivity extrapolates to a small residual value, so that the power extends over a range much larger than the residual resistivity and disorder may be considered unimportant. Our goal is to produce a model which produces a power law resistivity Tαsuperscript𝑇𝛼T^{\alpha}italic_T start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT with α<2𝛼2\alpha<2italic_α < 2 which is valid down to zero temperature and which include the linear resistivity as a special case α=1𝛼1\alpha=1italic_α = 1. As explained below (see also  [6]) there are barriers towards accomplishing this goal which may explain the paucity of such models. While our model is unlikely to be applicable to cuprate superconductors, it may find application in a doped organic compound. [9].

Why is it difficult to construct such a model? In the absence of disorder, the total momentum is conserved under scattering unless umklapp scattering is allowed. Therefore in the absence of both disorder and umklapp scattering the conductivity σ(ω)𝜎𝜔\sigma(\omega)italic_σ ( italic_ω ) is proportional to δ(ω)𝛿𝜔\delta(\omega)italic_δ ( italic_ω ) which is incompatible with any power law function ωαsuperscript𝜔𝛼\omega^{-\alpha}italic_ω start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT. Hence the coefficient of δ(ω)𝛿𝜔\delta(\omega)italic_δ ( italic_ω ) must be zero. This case was considered by Else and Senthil [10] who show that a power law resistivity requires the divergence of a certain kind of fluctuations in order to kill the delta function. We shall not pursue this route here and instead consider models where umklapp scattering is allowed down to T=0𝑇0T=0italic_T = 0. Such models immediately rule out scattering from critical modes at q=0𝑞0q=0italic_q = 0 such as those from ferromagnetic or nematic order or from emergent gauge fields. [6, 11] This leaves soft modes which are associated with critical point involving ordering at a finite momentum such as anti-ferromagnets or charge density waves. Such scatterings are limited to ”hot spots” on the Fermi surface. Without scatterings which rapidly equilibrate the other momentum states on the Fermi surface, the resistivity is dominated by scattering rates away from the hot spots. [12, 5] This problem can be brought under control with the introduction of disorder scattering. [5, 6, 11] This indeed gives rise to a linear T𝑇Titalic_T regime with a coefficient which is independent of the disorder. However, the price one pays is that this regime is limited to an increase of resistivity which is equal or less than the residual resistivity. Hence the zero disorder limit cannot be taken without losing the linear T𝑇Titalic_T regime. [5, 11]

It turns out that these difficulties can be circumvented in a model of do** holes into a quantum spin liquid with a spinon Fermi surface. The spinon Fermi surface has a singular self energy and violates the Landau criterion for quasi-particles. On a triangular lattice umklapp scattering is possible if the density of doped holes is large enough. We shall show that in this case the hot spot becomes a hot region on the Fermi surface, which may even cover the entire Fermi surface. So the bottleneck problem mentioned earlier does not arise and the power law resistivity survives to low temperature. We also emphasize that if the power law α𝛼\alphaitalic_α is equal to or less than unity, the Landau criterion is violated and quasi-particles as defined by Landau do not exist. Nevertheless, it was shown long ago by Prange and Kadanoff [13] that even without Landau’s quasi-particles, a Boltzmann equation that describes transport properties can be derived. Taking advantage of their insight, the resistivity can be calculated in a simple way even in the ”non-Fermi liquid” case. A quantum spin liquid with a spinon Fermi surface is an exotic state of matter, but it is no longer considered very strange. In fact, these state may be realized in certain organic compounds [14] and in monolayer 1T-TaSe22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT [15]. In any case, it is useful to have an example which can exhibit strange metal behavior based on a model that is not so strange.

Refer to caption
Figure 1: (a) Dashed circle shows the spinon Fermi surface in the hexagonal Brillouin zone for a triangular lattice. Solid blue circle shows the hole Fermi surface holding p𝑝pitalic_p doped holes. The two Fermi surfaces can exchange momentum 𝑷𝑷{\boldsymbol{P}}bold_italic_P via an umklapp process where 𝑷𝑷{\boldsymbol{P}}bold_italic_P is the sum of a 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT vector (dashed orange line) and a reciprocal lattice vector 𝑮𝑮{\boldsymbol{G}}bold_italic_G (green line) (b) An expanded view of the region between the two spinon Fermi surfaces shown in (a). 𝑷0subscript𝑷0{\boldsymbol{P}_{0}}bold_italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the shortest umklapp vector connecting the two spinon Fermi surfaces. The hole Fermi surface is positioned so that its Fermi surface is spanned by 𝑷0subscript𝑷0{\boldsymbol{P}_{0}}bold_italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. An initial hole state (green dot) at angle ϕ0subscriptitalic-ϕ0-\phi_{0}- italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is scattered to a final hole state (orange dot) at angle ϕ0+ϕsubscriptitalic-ϕ0italic-ϕ\phi_{0}+\phiitalic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_ϕ in an umklapp process. The short red arrow denote the vector Δ𝒒Δ𝒒\Delta{\boldsymbol{q}}roman_Δ bold_italic_q (defined above Eq. 2) which measures the deviation of the collective mode momentum 𝑸𝑸{\boldsymbol{Q}}bold_italic_Q from 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT

II The model

We consider a Hubbard model on the triangular lattice with nearest neighbor hop** t𝑡titalic_t and onsite U𝑈Uitalic_U. We assume U/t𝑈𝑡U/titalic_U / italic_t is such that we are on the insulator side of the Mott transition, but not deep inside, so the charge gap is finite but relatively small. S=1/2𝑆12S=1/2italic_S = 1 / 2 local moments are formed on the sites and we assume that they do not order but form a spin liquid state. We further assume that the electrons have fractionalized into fermionic spinons carrying S=1/2𝑆12S=1/2italic_S = 1 / 2 but no charge, and relativistic bosonic chargons. Both are coupled to emergent U(1) gauge fields. [16] This state was proposed to characterized the organic spin liquids compounds, the ET and dmit salts. [14] Recently the ET salt is found to have a spin gap below a phase transition at 6K, so if there is a Fermi surface, at leasst part of it is gapped out at low temperatures. [17] On the other hand the dmit salt does not show this transition or gap and continues to be a good candidate. For ET salts, we assume the spinon Fermi surface is a close-by competing state. More recently the spinon Fermi surface state was also proposed to be realized in monolayer 1T-TaSe22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT and 1T-TaS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT, and there is evidence of such as state from the appearance of incommensurate modulations at wave-vectors given by 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT expected in a Fermi liquid. [15] Further evidence comes from Kondo screening of adsorbed magnetic impurities [18] We note that recent DMRG calculations find a chiral spin liquid in the vicinity of the Mott transition and not a spinon Fermi surface. [19, 20] On the other hand, there is no experimental evidence for time reversal breaking in the two systems mentioned above. So we continue to assume that the Fermi surface state is either realized with additional coupling or is a competing state which may arise with carrier do**.

Now we consider do** this spin liquid with carriers which can be electrons or holes. For concreteness we shall use the hole notation. So far the 1T-TaSe22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT and 1T-TaS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT cannot be doped. However in the case of the ET salt, it has been possible to introduce Hg chains between the layers, resulting in a doped hole concentration p𝑝pitalic_p of 11 %. [9] At the mean-field level, the doped hole will occupy the gapped holon band. On the triangular lattice, the band minimum may occur at the ΓΓ\Gammaroman_Γ point or the K𝐾Kitalic_K points depending on the sign of t𝑡titalic_t. Here we make a further assumption. We consider a strong attraction between the holon and the spinon so they recombine to form physical holes that carry both charge and spin. These holes form a Fermi surface containing p𝑝pitalic_p holes if the band bottom is located at ΓΓ\Gammaroman_Γ, or they form two Fermi surfaces containing p/2𝑝2p/2italic_p / 2 holes each, if the band bottoms are at the zone corner K𝐾Kitalic_K points. We shall refer to such bands as hole bands. Alternatively, there may be an additional band which happens to be located below the chargon gap, and the doped holes enter that band to form conventional hole pockets. There is in fact evidence for such a band in 1T-TaSe22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT [21]. The situation is illustrated in Fig.1. The spinon band is half filled and is almost circular. For concreteness the hole band is assumed to be at the ΓΓ\Gammaroman_Γ point and has area corresponding to p𝑝pitalic_p spinful holes.

Note that the hole band containing p𝑝pitalic_p holes violates Luttinger theorem which states that the Fermi surfaces should enclose a total of 1p1𝑝1-p1 - italic_p fermions. This situation has been dubbed FL*{}^{*}start_FLOATSUPERSCRIPT * end_FLOATSUPERSCRIPT by Senthil, Sachdev and Vojta. [22] They explained that one can get around the non-perturbative proof of Luttinger theorem [23] if the spinon sector is in a gapped topological state with degenerate ground states upon flux insertion in a torus geometry. In our case, the spinon sector is gapless and has massive degeneracy. As a result the non-perturbative proof which relies on returning to the same ground state after a flux insertion in a torus also fails. [24]. While topology does not play a role, we follow them and denote this as an FL*{}^{*}start_FLOATSUPERSCRIPT * end_FLOATSUPERSCRIPT state.

Refer to caption
Figure 2: A plot of the exponent α𝛼\alphaitalic_α which characterizes the resistivity going as Tαsuperscript𝑇𝛼T^{\alpha}italic_T start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT vs the exponent σ𝜎\sigmaitalic_σ. Dashed line indicates the breakdown of the linear relation given by Eq.5 for σ>5/6𝜎56\sigma>5/6italic_σ > 5 / 6. The Fermi liquid behave α=2𝛼2\alpha=2italic_α = 2 is recovered for σ<1/6𝜎16\sigma<1/6italic_σ < 1 / 6 and linear T resistivity obtains for σ=2/3𝜎23\sigma=2/3italic_σ = 2 / 3, as indicated by the short-dashed line. σ𝜎\sigmaitalic_σ is defined as the divergent exponent of the vertex function γ(ω)ωσproportional-to𝛾𝜔superscript𝜔𝜎\gamma(\omega)\propto\omega^{-\sigma}italic_γ ( italic_ω ) ∝ italic_ω start_POSTSUPERSCRIPT - italic_σ end_POSTSUPERSCRIPT. As shown in the inset this divergence is due to repeated exchange of the gauge field (dotted line). Also shown is the diagram for the spinon polarization function Π2kFsubscriptΠ2subscript𝑘𝐹\Pi_{2k_{F}}roman_Π start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT. The intermediate spinon lines denote excitations of spinon particle-hole pairs which are restricted to low energy, hence two factors of the vertex functions appear and contribute the factor ω~2σsuperscript~𝜔2𝜎\tilde{\omega}^{-2\sigma}over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT - 2 italic_σ end_POSTSUPERSCRIPT in Eq. 1. For details see ref. [25]

The properties of the hole pocket can be quite conventional, apart from its small Fermi surface area. In contrast, the spinon Fermi surface is strongly coupled to a U(1) gauge field which the holes do not see. This problem has been well studied and one key result is that the self energy goes as ω2/3superscript𝜔23\omega^{2/3}italic_ω start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT[26] The spinon decay rate exceed ω𝜔\omegaitalic_ω which violates Landau’s criterion for the existence of quasi-particles. The origin of the strong decay rate is that the gauge field is strongly overdamped, with a propagator that goes as (|ω|/q+q2)1superscript𝜔𝑞superscript𝑞21(|\omega|/q+q^{2})^{-1}( | italic_ω | / italic_q + italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The excitation ω𝜔\omegaitalic_ω scales as q3superscript𝑞3q^{3}italic_q start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resulting in a copious amount of low energy excitations. For a number of years it was believed that this strong coupled problem can be controlled with a 1/N1𝑁1/N1 / italic_N expansion, where N𝑁Nitalic_N is the number of fermion species. The low energy physics is found to be described by a nontrivial fixed point which is exactly marginal and described by scaling functions. [25, 27] Of particular interest to us is that the polarizibility function for the system comprising spinons coupled to gauge fields, Π(ω,𝑸)Π𝜔𝑸\Pi(\omega,{\boldsymbol{Q}})roman_Π ( italic_ω , bold_italic_Q ), is singular for |𝑸|𝑸|{\boldsymbol{Q}}|| bold_italic_Q | near 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT[25] Importantly, note that this results in a ring of low energy excitations in momentum space with radius 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT. Unfortunately, upon further scrutiny it was shown by Sung-Sik Lee that the 1/N1𝑁1/N1 / italic_N is not controlled even in the large N𝑁Nitalic_N limit [28]. So the nature of the low energy physics is not known even for large N𝑁Nitalic_N, let alone the physical case of N=2𝑁2N=2italic_N = 2. Nevertheless, it is still possible that the low energy physics remains to be described by a nontrivial and marginal fixed point. There is support for this possibility by studies involving more expansion parameters. [29] We will adopt this point of view and assume a scaling form for Π(ω,𝑸)Π𝜔𝑸\Pi(\omega,{\boldsymbol{Q}})roman_Π ( italic_ω , bold_italic_Q ) which is the same for the large N𝑁Nitalic_N expansion as given by Altshuler, Ioffe and Millis [25], except that we will treat the scaling exponent σ𝜎\sigmaitalic_σ as an unknown parameter. (σ𝜎\sigmaitalic_σ is the exponent which characterize the divergence of the 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT vertex function γ𝛾\gammaitalic_γ due to coupling to the gauge fluctuations as shown in Fig 2, namely, γ(ω)|ω|σproportional-to𝛾𝜔superscript𝜔𝜎\gamma(\omega)\propto|\omega|^{-\sigma}italic_γ ( italic_ω ) ∝ | italic_ω | start_POSTSUPERSCRIPT - italic_σ end_POSTSUPERSCRIPT.) Since Π(ω,𝑸)Π𝜔𝑸\Pi(\omega,{\boldsymbol{Q}})roman_Π ( italic_ω , bold_italic_Q ) is independent of the direction of 𝑸𝑸{\boldsymbol{Q}}bold_italic_Q, we introduce the variable Δ𝒒=𝑸2kF𝑸^Δ𝒒𝑸2subscript𝑘𝐹^𝑸\Delta{\boldsymbol{q}}={\boldsymbol{Q}}-2k_{F}\hat{{\boldsymbol{Q}}}roman_Δ bold_italic_q = bold_italic_Q - 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT over^ start_ARG bold_italic_Q end_ARG to denote the distance of 𝑸𝑸{\boldsymbol{Q}}bold_italic_Q from the Fermi surface. We write Π(ω,𝑸)Π𝜔𝑸\Pi(\omega,{\boldsymbol{Q}})roman_Π ( italic_ω , bold_italic_Q ) as Π2kF(ω,Δ𝒒)subscriptΠ2subscript𝑘𝐹𝜔Δ𝒒\Pi_{2k_{F}}(\omega,\Delta{\boldsymbol{q}})roman_Π start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_ω , roman_Δ bold_italic_q ) and denote the imaginary part as Π′′superscriptΠ′′\Pi^{{}^{\prime\prime}}roman_Π start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT. We define the scaled variable q~~𝑞\tilde{q}over~ start_ARG italic_q end_ARG as q~𝑸^=Δ𝒒/kF~𝑞^𝑸Δ𝒒subscript𝑘𝐹\tilde{q}\hat{{\boldsymbol{Q}}}=\Delta{\boldsymbol{q}}/k_{F}over~ start_ARG italic_q end_ARG over^ start_ARG bold_italic_Q end_ARG = roman_Δ bold_italic_q / italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, together with ω~=ω/EF~𝜔𝜔subscript𝐸𝐹\tilde{\omega}=\omega/E_{F}over~ start_ARG italic_ω end_ARG = italic_ω / italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT. We have,

Π2kF′′(ω,Δ𝒒)1EFω~2/32σ,ω~2/3>|q~|formulae-sequenceproportional-tosubscriptsuperscriptΠ′′2subscript𝑘𝐹𝜔Δ𝒒1subscript𝐸𝐹superscript~𝜔232𝜎superscript~𝜔23~𝑞\Pi^{{}^{\prime\prime}}_{2k_{F}}(\omega,\Delta{\boldsymbol{q}})\propto\frac{1}% {E_{F}}\tilde{\omega}^{2/3-2\sigma},\quad\tilde{\omega}^{2/3}>|\tilde{q}|roman_Π start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_ω , roman_Δ bold_italic_q ) ∝ divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 - 2 italic_σ end_POSTSUPERSCRIPT , over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT > | over~ start_ARG italic_q end_ARG | (1)
Π2kF′′(ω,Δ𝒒){1EFω~|q~|1/23σ,ω~2/3<|q~|,q~<0.1EFω~5/3|q~|3/23σ,ω~2/3<|q~|,q~>0.proportional-tosubscriptsuperscriptΠ′′2subscript𝑘𝐹𝜔Δ𝒒casesformulae-sequence1subscript𝐸𝐹~𝜔superscript~𝑞123𝜎superscript~𝜔23~𝑞~𝑞0𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒formulae-sequence1subscript𝐸𝐹superscript~𝜔53superscript~𝑞323𝜎superscript~𝜔23~𝑞~𝑞0𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒\Pi^{{}^{\prime\prime}}_{2k_{F}}(\omega,\Delta{\boldsymbol{q}})\propto\begin{% cases}\frac{1}{E_{F}}\tilde{\omega}|\tilde{q}|^{-1/2-3\sigma},\quad\tilde{% \omega}^{2/3}<|\tilde{q}|,\tilde{q}<0.\\ \frac{1}{E_{F}}\tilde{\omega}^{5/3}|\tilde{q}|^{-3/2-3\sigma},\quad\tilde{% \omega}^{2/3}<|\tilde{q}|,\tilde{q}>0.\par\end{cases}roman_Π start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_ω , roman_Δ bold_italic_q ) ∝ { start_ROW start_CELL divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG over~ start_ARG italic_ω end_ARG | over~ start_ARG italic_q end_ARG | start_POSTSUPERSCRIPT - 1 / 2 - 3 italic_σ end_POSTSUPERSCRIPT , over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT < | over~ start_ARG italic_q end_ARG | , over~ start_ARG italic_q end_ARG < 0 . end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 5 / 3 end_POSTSUPERSCRIPT | over~ start_ARG italic_q end_ARG | start_POSTSUPERSCRIPT - 3 / 2 - 3 italic_σ end_POSTSUPERSCRIPT , over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT < | over~ start_ARG italic_q end_ARG | , over~ start_ARG italic_q end_ARG > 0 . end_CELL start_CELL end_CELL end_ROW (2)

In these equations, the σ𝜎\sigmaitalic_σ dependent part comes from the vertex function γ𝛾\gammaitalic_γ and the rest comes from particle-hole excitations including the self-energy correction. Eq.1 gives the limit ω~2/3>|q~|superscript~𝜔23~𝑞\tilde{\omega}^{2/3}>|\tilde{q}|over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT > | over~ start_ARG italic_q end_ARG | and is given in ref. [25]. The limit ω~2/3<|q~|superscript~𝜔23~𝑞\tilde{\omega}^{2/3}<|\tilde{q}|over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT < | over~ start_ARG italic_q end_ARG | is given in Eq.2. The first line is applicable for q~<0~𝑞0\tilde{q}<0over~ start_ARG italic_q end_ARG < 0 or |𝑸|<2kF𝑸2subscript𝑘𝐹|{\boldsymbol{Q}}|<2k_{F}| bold_italic_Q | < 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT and apart from the σ𝜎\sigmaitalic_σ dependent factor, it is the same as the familiar form for free fermions. [30] The only difference is that the condition of validity is changed from ω~<|q~|~𝜔~𝑞\tilde{\omega}<|\tilde{q}|over~ start_ARG italic_ω end_ARG < | over~ start_ARG italic_q end_ARG | to ω~2/3<|q~|superscript~𝜔23~𝑞\tilde{\omega}^{2/3}<|\tilde{q}|over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT < | over~ start_ARG italic_q end_ARG |. The second line in Eq.2 gives the case |𝑸|>2kF𝑸2subscript𝑘𝐹|{\boldsymbol{Q}}|>2k_{F}| bold_italic_Q | > 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT which is zero for free fermions because we are outside of the particle-hole continuum. In our case there is a finite contribution due to a self energy which goes as ω2/3superscript𝜔23\omega^{2/3}italic_ω start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT. Note that Eqs.1 and 2 satisfy the scaling form Π2kF′′ω~2/32σF(ω~/q~3/2)proportional-tosuperscriptsubscriptΠ2subscript𝑘𝐹′′superscript~𝜔232𝜎𝐹~𝜔superscript~𝑞32\Pi_{2k_{F}}^{{}^{\prime\prime}}\propto\tilde{\omega}^{2/3-2\sigma}F(\tilde{% \omega}/\tilde{q}^{3/2})roman_Π start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ∝ over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 - 2 italic_σ end_POSTSUPERSCRIPT italic_F ( over~ start_ARG italic_ω end_ARG / over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT ) where F(x)𝐹𝑥F(x)italic_F ( italic_x ) goes to 1 for x𝑥xitalic_x small so that ω~~𝜔\tilde{\omega}over~ start_ARG italic_ω end_ARG scales as q~3/2superscript~𝑞32\tilde{q}^{3/2}over~ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT.

We note that while Eqs.1 and 2 are often derived assuming a circular Fermi surface, it is generally applicable to any Fermi surface shape as long as opposite k𝑘kitalic_k points on the Fermi surface have parallel tangents. In this case kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is a function of angle θ𝜃\thetaitalic_θ and Eqs.1 and 2 remain valid.

III Resistivity

Now we are ready to compute the resistivity of the hole band due to scattering by the soft 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT mode of the spinon Fermi surface. We note that after shifting by a reciprocal lattice vector 𝑮𝑮{\boldsymbol{G}}bold_italic_G, the 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT vectors are equivalent to a set of vectors 𝑷𝑷{\boldsymbol{P}}bold_italic_P centered at the M point that connect the Fermi surfaces on neigboring extended Brillouin zones, as shown in fig 1a. If the hole Fermi surface is large enough, these vector can connect points on the hole Fermi surface and give rise to umklapp scattering which relaxes the momentum and current. The condition on the size of the hole Fermi surface is the following. With nearest neighbor hop** the spinon Fermi surface is nearly a circle with radius kF0.375|𝒃1|subscript𝑘𝐹0.375subscript𝒃1k_{F}\approx 0.375|{\boldsymbol{b}_{1}}|italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ≈ 0.375 | bold_italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | where 𝒃1subscript𝒃1{\boldsymbol{b}_{1}}bold_italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the reciprocal lattice vector along x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG[15] The length of the shortest vector that connects that two Fermi surfaces is 0.25|𝒃1|0.25subscript𝒃10.25|{\boldsymbol{b}_{1}}|0.25 | bold_italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT |. Let us denote the Fermi momentum of the hole pocket by pFsubscript𝑝𝐹p_{F}italic_p start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT. The condition for umklapp is that pF>0.125|𝒃1|subscript𝑝𝐹0.125subscript𝒃1p_{F}>0.125|{\boldsymbol{b}_{1}}|italic_p start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT > 0.125 | bold_italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | which is 1/3 of kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT for 1 spinon per unit cell. Hence we conclude that umklapp scattering begins for p>1/9𝑝19p>1/9italic_p > 1 / 9 in the case of a single circular hole pocket centered at ΓΓ\Gammaroman_Γ. This condition can be relaxed if the Fermi surface deviates from a circle, which is likely the case if the doped holes occupy a separate trivial band. Below we will assume that this condition is satisfied and calculate the hole lifetime due to umklapp scattering. Since the momentum transfer is large and umklapp, the same lifetime will enter the resistivity and all transport phenomena.

Unlike the spinon Fermi surface, the hole Fermi surface is not coupled to the gauge field and does not have the anomalous large self energy and dam** rate. The major source of dam** at low temperatures come from the umklapp scattering channel under consideration. If this rate is smaller than linear in T𝑇Titalic_T or its frequency ΩΩ\Omegaroman_Ω, the quasi-particle is well defined and a Fermi liquid description is valid in the Landau sense, even if Luttinger theorem is not obeyed. Here I remark that even if the decay rate ends up with a power law smaller than unity and Landau quasi-particles do not exist, it is still possible to treat the transport problem using Boltzmann equation, as long as the self energy has only frequency dependence and no singular momentum dependence, which is the case here. This was shown by Prange and Kadanoff [13] and their idea have been applied to the fermion coupled to gauge field problem. [31, 11] The idea is that at low frequency, the electron spectral function is a sharp peak in momentum space, and a Fermi surface can be defined by the crossing of this sharp peak across pFsubscript𝑝𝐹p_{F}italic_p start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT. This is familiar in the angle resolved photo-emission spectroscopy (ARPES) literature in that the momentum distribution curve (MDC) can be sharp while the energy distribution curve (EDC) is broad. This allows us to use the Boltzmann equation approach to calculate the resistivity and the result will remain valid even if the exponent α𝛼\alphaitalic_α ends up being less than unity, which will happen in certain parameter range.

In its simplest form, the solution of the Boltzmann equation is just the computation of the scattering rate 1/τ1𝜏1/\tau1 / italic_τ using Fermi’s golden rule. As shown in Fig.1b, we label the state on the hole Fermi surface by an angle ϕitalic-ϕ\phiitalic_ϕ. For simplicity of exposition, we consider the scattering from an initial state at ϕ0subscriptitalic-ϕ0-\phi_{0}- italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to a final state at the Fermi surface near ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT such that they are connected by the shortest umklapp vector which lies along x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG. 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT scattering using longer spanning wave-vector will have similar scattering rate with the same power law. The important point is to note that in general a finite region of ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values will satisfy this umklapp condition and will have similar umklapp scattering rates. Furthermore there are 6 minimal spanning vectors in total which replicate the one shown in fig 1b. Therefore it is quite likely that these regions of initial states will cover the entire Fermi surface. This situation is totally different from the scattering from a critical mode at a finite momentum, such as that due to antiferromagnetic instability. This results in the so called ”hot spots” and the problem is that the rapid relaxation is limited to these hot spots and there is a bottleneck to relax momenta from the rest of the Fermi surface. [12, 5] In order to overcome the bottleneck constraint, one need to introduce disorder scattering which gives a finite resistivity at zero temperature, leading to the difficulties described earlier. An important feature of the current model is that the entire Fermi surface or large fragments of it is hot, which allows us to reach the clean limit.

The umklapp scattering rate is given by [11]

1/τ=V020Ω𝑑ω𝑑ϕΠ2kF"(ω,Δ𝒒(ϕ))1𝜏superscriptsubscript𝑉02superscriptsubscript0Ωdifferential-d𝜔differential-ditalic-ϕsuperscriptsubscriptΠ2subscript𝑘𝐹"𝜔Δ𝒒italic-ϕ1/\tau=V_{0}^{2}\int_{0}^{\Omega}d\omega\int d\phi\ \Pi_{2k_{F}}^{"}(\omega,% \Delta{\boldsymbol{q}(\phi)})1 / italic_τ = italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Ω end_POSTSUPERSCRIPT italic_d italic_ω ∫ italic_d italic_ϕ roman_Π start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT " end_POSTSUPERSCRIPT ( italic_ω , roman_Δ bold_italic_q ( italic_ϕ ) ) (3)

where V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a short range interaction constant between the holes and the spinons. Since the imaginary part of Π2kFsubscriptΠ2subscript𝑘𝐹\Pi_{2k_{F}}roman_Π start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT represents the excitation of a particle-hole pair of spinons, this equation captures the scattering of holes by spinons described in Fig.1. The integral is over a final state located at ϕ0+ϕsubscriptitalic-ϕ0italic-ϕ\phi_{0}+\phiitalic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_ϕ on the Fermi surface and 𝑸(ϕ)𝑸italic-ϕ{\boldsymbol{Q}(\phi)}bold_italic_Q ( italic_ϕ ) denotes a vector in the direction connecting this point and the center of the spinon Fermi surface circle which is close to x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG for small ϕitalic-ϕ\phiitalic_ϕ. We are interested in 𝑸(ϕ)𝑸italic-ϕ{\boldsymbol{Q}}(\phi)bold_italic_Q ( italic_ϕ ) near 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, so the integration over ϕitalic-ϕ\phiitalic_ϕ can be converted to a one dimensional integral over the length of Δ𝒒Δ𝒒\Delta{\boldsymbol{q}}roman_Δ bold_italic_q where Δ𝒒Δ𝒒\Delta{\boldsymbol{q}}roman_Δ bold_italic_q is the deviation of 𝑸𝑸{\boldsymbol{Q}}bold_italic_Q from the 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT vector as defined earlier. It is easy to see that |Δ𝒒|kFϕcos(ϕ0)Δ𝒒subscript𝑘𝐹italic-ϕ𝑐𝑜𝑠subscriptitalic-ϕ0|\Delta{\boldsymbol{q}}|\approx k_{F}\phi cos(\phi_{0})| roman_Δ bold_italic_q | ≈ italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_ϕ italic_c italic_o italic_s ( italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ). Hence the integral over ϕitalic-ϕ\phiitalic_ϕ is converted to an integral over q~/cos(ϕ0)~𝑞𝑐𝑜𝑠subscriptitalic-ϕ0\tilde{q}/cos(\phi_{0})over~ start_ARG italic_q end_ARG / italic_c italic_o italic_s ( italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ). We have

1/τ=V02cos(ϕ0)0Ω𝑑ωΛΛ𝑑q~Π2kF"(ω,Δ𝒒(ϕ))1𝜏superscriptsubscript𝑉02𝑐𝑜𝑠subscriptitalic-ϕ0superscriptsubscript0Ωdifferential-d𝜔superscriptsubscriptΛΛdifferential-d~𝑞superscriptsubscriptΠ2subscript𝑘𝐹"𝜔Δ𝒒italic-ϕ1/\tau=\frac{V_{0}^{2}}{cos(\phi_{0})}\int_{0}^{\Omega}d\omega\int_{-\Lambda}^% {\Lambda}d\tilde{q}\ \Pi_{2k_{F}}^{"}(\omega,\Delta{\boldsymbol{q}(\phi)})1 / italic_τ = divide start_ARG italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_c italic_o italic_s ( italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Ω end_POSTSUPERSCRIPT italic_d italic_ω ∫ start_POSTSUBSCRIPT - roman_Λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Λ end_POSTSUPERSCRIPT italic_d over~ start_ARG italic_q end_ARG roman_Π start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT " end_POSTSUPERSCRIPT ( italic_ω , roman_Δ bold_italic_q ( italic_ϕ ) ) (4)

where Λ1Λ1\Lambda\approx 1roman_Λ ≈ 1 is an ultra-violet (UV) cut-off in the q~~𝑞\tilde{q}over~ start_ARG italic_q end_ARG integration. We can divide the q~~𝑞\tilde{q}over~ start_ARG italic_q end_ARG integral into two regions. Region (1) is for ω~2/3>|q~|superscript~𝜔23~𝑞\tilde{\omega}^{2/3}>|\tilde{q}|over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT > | over~ start_ARG italic_q end_ARG | as given by Eq.1. The q~~𝑞\tilde{q}over~ start_ARG italic_q end_ARG integral gives a factor ω~2/3superscript~𝜔23\tilde{\omega}^{2/3}over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT and we find 1/τΩ7/32σproportional-to1𝜏superscriptΩ732𝜎1/\tau\propto\Omega^{7/3-2\sigma}1 / italic_τ ∝ roman_Ω start_POSTSUPERSCRIPT 7 / 3 - 2 italic_σ end_POSTSUPERSCRIPT Region (2) is for ω~2/3<|q~|superscript~𝜔23~𝑞\tilde{\omega}^{2/3}<|\tilde{q}|over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT < | over~ start_ARG italic_q end_ARG | as given by Eq.2. We consider separately the contributions for |𝑸|𝑸|{\boldsymbol{Q}}|| bold_italic_Q | less than or greater than 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT . In the latter case the q𝑞{q}italic_q integral is UV convergent since σ𝜎\sigmaitalic_σ is positive, so its value is given by the infra-red cut-off and we find a contribution equal to that of region 1. On the other hand, for |𝑸|<2kF𝑸2subscript𝑘𝐹|{\boldsymbol{Q}}|<2k_{F}| bold_italic_Q | < 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT the integrand is given by the first line in Eq.2 we see that the integral is UV convergent only for σ>1/6𝜎16\sigma>1/6italic_σ > 1 / 6. Hence we find the exponent for the scattering rate 1/τΩαproportional-to1𝜏superscriptΩ𝛼1/\tau\propto\Omega^{\alpha}1 / italic_τ ∝ roman_Ω start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT or Tαsuperscript𝑇𝛼T^{\alpha}italic_T start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT, where

α=7/32σ,σ>1/6formulae-sequence𝛼732𝜎𝜎16\alpha=7/3-2\sigma,\quad\sigma>1/6italic_α = 7 / 3 - 2 italic_σ , italic_σ > 1 / 6 (5)

Note that this result is a consequence of the scaling property for Π2kFsubscriptΠ2subscript𝑘𝐹\Pi_{2k_{F}}roman_Π start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT described earlier following Eq.2 and does not depend on the values of the exponents which only serve to control the regime of validity as we next discuss. From the first line in Eq.2, for σ<1/6𝜎16\sigma<1/6italic_σ < 1 / 6, the q~~𝑞\tilde{q}over~ start_ARG italic_q end_ARG integral is infra-red convergent, so its value is given by the UV cutoff ΛΛ\Lambdaroman_Λ and is independent of ω~~𝜔\tilde{\omega}over~ start_ARG italic_ω end_ARG. As a result, we find that α=2𝛼2\alpha=2italic_α = 2 which dominates over contributions from other regions of the integratioan and we recover the standard Landau Fermi liquid result. On the other hand the spinon self energy is given by ωαsuperscript𝜔𝛼\omega^{\alpha}italic_ω start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT due to scattering by the holes. When α<2/3𝛼23\alpha<2/3italic_α < 2 / 3 the exponent will need to be determined self-consistently. Hence the the validity of Eq.5 is limited to the range 2/3<α<7/323𝛼732/3<\alpha<7/32 / 3 < italic_α < 7 / 3. This final result is summarized in Fig 2.

We do not know what the value of σ𝜎\sigmaitalic_σ is, but as a matter of principle, this model gives a power law behavior of the resistivity down to the lowest temperature which can include the linear T resistivity. In fact, the linear T case α=1𝛼1\alpha=1italic_α = 1 does not play any special role and there is no obvious restriction that α𝛼\alphaitalic_α cannot be smaller than unity.

IV Discussion

We have constructed a model based on do** of a spin liquid with a spinon Fermi surface, forming a hole Fermi surface with an area that corresponds to the doped hole concentration p𝑝pitalic_p. As such it violates Luttinger theorem and belongs to the classification FL*𝐹superscript𝐿FL^{*}italic_F italic_L start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT[22] We show that for p𝑝pitalic_p large enough, umklapp scattering between the hole Fermi surface and the spinon Fermi surface becomes possible, resulting in a resistivity which goes as Tαsuperscript𝑇𝛼T^{\alpha}italic_T start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT where α𝛼\alphaitalic_α is given by Eq.5 and can vary over a wide range, depending on the strength of the 2kF2subscript𝑘𝐹2k_{F}2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT singularity of the spinon. In principle α𝛼\alphaitalic_α equal to unity or even smaller in possible. In these cases, the Landau criterion for the existence of his quasi-particle is violated and this state can be called a non-Fermi liquid. Our calculation relies on the formulation of Prange and Kadanoff  [13] who showed that quasi-particles in the Landau sense is not necessary to derive a Boltzmann equation to describe transport. Therefore our result should remain valid for α𝛼\alphaitalic_α equal to or less than unity. On the other hand, since the transport properties are based on the Boltzmann equation, other properties such as the Hall constant and magnetoresistance should be conventional. In particular, the magnetoresistance should obey Koehler’s rule, which states that the correction to resistivity goes as (Bτ)2superscript𝐵𝜏2(B\tau)^{2}( italic_B italic_τ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. In cuprates a linear in B𝐵Bitalic_B magnetoresistance is often associated with a linear in T𝑇Titalic_T resistivity. It is unlikely that our model or something similar is relevant to the cuprates. The most promising material candidate is the doped organic system where 11% do** has been achieved and linear or close to linear in T resistivity which extends over several times of the residual resistivity have been observed over some range of pressure. [9] The Hall constant is given by the hole do** in this pressure range, consistent with small Fermi pockets of total area p𝑝pitalic_p. Interestingly, 11% is on the border of applicability of our model, which requires p>1/9𝑝19p>1/9italic_p > 1 / 9.

We note that the key ingredient of our model is a critical mode which is soft along a line in momentum space. This gives rise to low energy scattering in finite regions on the Fermi surface which allows us to circumvent the hot spot problem associated with scattering by a mode which is critical at one momentum. We employ a model with a spinon Fermi surface because that represents a critical state which exists over a range in parameter space. Therefore the power law resistivity exists over a range of parameters such as do** or pressure. Furthermore a circular spinon or hole Fermi surface is not needed, because the singularity of Π2kFsubscriptΠ2subscript𝑘𝐹\Pi_{2k_{F}}roman_Π start_POSTSUBSCRIPT 2 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_POSTSUBSCRIPT is a consequence of a two patch model, which rely on the fact that opposite k points in the spinon Fermi surface have parallel tangents.

There has been a lot of discussions concerning the so-called Planckian bound, /τ<kBTPlanck-constant-over-2-pi𝜏subscript𝑘𝐵𝑇\hbar/\tau<k_{B}Troman_ℏ / italic_τ < italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T[32] At present there appears to be no clear connection between transport lifetime and the Planckian bound. [2] Our model is relatively simple and can serve as a counter-example to the Plankian bound which, unlike phonon scattering, is valid down to zero temperature.

Acknowledgement

I thank Andrey Chubukov and T. Senthil for enlightening discussions and acknowledges support by DOE (USA) office of Basic Sciences Grant No. DE-FG02-03ER46076.

References

  • Phillips et al. [2022] P. W. Phillips, N. E. Hussey, and P. Abbamonte, Stranger than metals, Science 377, eabh4273 (2022).
  • Hartnoll and Mackenzie [2022] S. A. Hartnoll and A. P. Mackenzie, Colloquium: Planckian dissipation in metals, Reviews of Modern Physics 94, 041002 (2022).
  • Putzke et al. [2021] C. Putzke, S. Benhabib, W. Tabis, J. Ayres, Z. Wang, L. Malone, S. Licciardello, J. Lu, T. Kondo, T. Takeuchi, et al., Reduced hall carrier density in the overdoped strange metal regime of cuprate superconductors, Nature Physics 17, 826 (2021).
  • Greene et al. [2020] R. L. Greene, P. R. Mandal, N. R. Poniatowski, and T. Sarkar, The strange metal state of the electron-doped cuprates, Annual Review of Condensed Matter Physics 11, 213 (2020).
  • Rosch [2000] A. Rosch, Magnetotransport in nearly antiferromagnetic metals, Physical Review B 62, 4945 (2000).
  • Maslov et al. [2011] D. L. Maslov, V. I. Yudson, and A. V. Chubukov, Resistivity of a non-galilean–invariant fermi liquid near pomeranchuk quantum criticality, Physical review letters 106, 106403 (2011).
  • Cha et al. [2020] P. Cha, N. Wentzell, O. Parcollet, A. Georges, and E.-A. Kim, Linear resistivity and sachdev-ye-kitaev (syk) spin liquid behavior in a quantum critical metal with spin-1/2 fermions, Proceedings of the National Academy of Sciences 117, 18341 (2020).
  • Patel et al. [2023] A. A. Patel, H. Guo, I. Esterlis, and S. Sachdev, Universal theory of strange metals from spatially random interactions, Science 381, 790 (2023).
  • Suzuki et al. [2022] Y. Suzuki, K. Wakamatsu, J. Ibuka, H. Oike, T. Fujii, K. Miyagawa, H. Taniguchi, and K. Kanoda, Mott-driven bec-bcs crossover in a doped spin liquid candidate κ𝜅\kappaitalic_κ-(bedt- ttf) 4 hg 2.89 br 8, Physical Review X 12, 011016 (2022).
  • Else and Senthil [2021] D. V. Else and T. Senthil, Strange metals as ersatz fermi liquids, Physical Review Letters 127, 086601 (2021).
  • Lee [2021] P. A. Lee, Low-temperature t-linear resistivity due to umklapp scattering from a critical mode, Physical Review B 104, 035140 (2021).
  • Hlubina and Rice [1995] R. Hlubina and T. Rice, Resistivity as a function of temperature for models with hot spots on the fermi surface, Physical Review B 51, 9253 (1995).
  • Prange and Kadanoff [1964] R. E. Prange and L. P. Kadanoff, Transport theory for electron-phonon interactions in metals, Physical Review 134, A566 (1964).
  • Zhou et al. [2017] Y. Zhou, K. Kanoda, and T.-K. Ng, Quantum spin liquid states, Reviews of Modern Physics 89, 025003 (2017).
  • Ruan et al. [2021] W. Ruan, Y. Chen, S. Tang, J. Hwang, H.-Z. Tsai, R. L. Lee, M. Wu, H. Ryu, S. Kahn, F. Liou, et al., Evidence for quantum spin liquid behaviour in single-layer 1t-tase2 from scanning tunnelling microscopy, Nature Physics 17, 1154 (2021).
  • Lee and Lee [2005] S.-S. Lee and P. A. Lee, U (1) gauge theory of the hubbard model: Spin liquid states and possible application to κ𝜅\kappaitalic_κ-(bedt- ttf) 2 cu 2 (cn) 3, Physical review letters 95, 036403 (2005).
  • Miksch et al. [2021] B. Miksch, A. Pustogow, M. J. Rahim, A. A. Bardin, K. Kanoda, J. A. Schlueter, R. Hübner, M. Scheffler, and M. Dressel, Gapped magnetic ground state in quantum spin liquid candidate κ𝜅\kappaitalic_κ-(bedt-ttf) 2cu2 (cn) 3, Science 372, 276 (2021).
  • Chen et al. [2022a] Y. Chen, W.-Y. He, W. Ruan, J. Hwang, S. Tang, R. L. Lee, M. Wu, T. Zhu, C. Zhang, H. Ryu, et al., Evidence for a spinon kondo effect in cobalt atoms on single-layer 1t-tase2, Nature Physics 18, 1335 (2022a).
  • Szasz et al. [2020] A. Szasz, J. Motruk, M. P. Zaletel, and J. E. Moore, Chiral spin liquid phase of the triangular lattice hubbard model: a density matrix renormalization group study, Physical Review X 10, 021042 (2020).
  • Chen et al. [2022b] B.-B. Chen, Z. Chen, S.-S. Gong, D. Sheng, W. Li, and A. Weichselbaum, Quantum spin liquid with emergent chiral order in the triangular-lattice hubbard model, Physical Review B 106, 094420 (2022b).
  • Chen et al. [2020] Y. Chen, W. Ruan, M. Wu, S. Tang, H. Ryu, H.-Z. Tsai, R. L. Lee, S. Kahn, F. Liou, C. Jia, et al., Strong correlations and orbital texture in single-layer 1t-tase2, Nature Physics 16, 218 (2020).
  • Senthil et al. [2003] T. Senthil, S. Sachdev, and M. Vojta, Fractionalized fermi liquids, Physical review letters 90, 216403 (2003).
  • Oshikawa [2000] M. Oshikawa, Topological approach to luttinger’s theorem and the fermi surface of a kondo lattice, Physical Review Letters 84, 3370 (2000).
  • Senthil et al. [2004] T. Senthil, M. Vojta, and S. Sachdev, Weak magnetism and non-fermi liquids near heavy-fermion critical points, Physical Review B 69, 035111 (2004).
  • Altshuler et al. [1994] B. Altshuler, L. Ioffe, and A. Millis, Low-energy properties of fermions with singular interactions, Physical Review B 50, 14048 (1994).
  • Lee and Nagaosa [1992] P. A. Lee and N. Nagaosa, Gauge theory of the normal state of high-t c superconductors, Physical Review B 46, 5621 (1992).
  • Polchinski [1994] J. Polchinski, Low-energy dynamics of the spinon-gauge system, Nuclear Physics B 422, 617 (1994).
  • Lee [2009] S.-S. Lee, Low-energy effective theory of fermi surface coupled with u (1) gauge field in 2+ 1 dimensions, Physical Review B 80, 165102 (2009).
  • Mross et al. [2010] D. F. Mross, J. McGreevy, H. Liu, and T. Senthil, Controlled expansion for certain non-fermi-liquid metals, Physical Review B 82, 045121 (2010).
  • Zhang et al. [2023] S.-S. Zhang, E. Berg, and A. V. Chubukov, Free energy and specific heat near a quantum critical point of a metal, Physical Review B 107, 144507 (2023).
  • Kim et al. [1995] Y. B. Kim, P. A. Lee, X.-G. Wen, and P. Stamp, Influence of gauge-field fluctuations on composite fermions near the half-filled state, Physical Review B 51, 10779 (1995).
  • Zaanen [2004] J. Zaanen, Why the temperature is high, Nature 430, 512 (2004).