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

Anisotropic scattering caused by apical oxygen vacancies in thin films of overdoped high-temperature cuprate superconductors

Da Wang [email protected]    Jun-Qi Xu    Hai-Jun Zhang    Qiang-Hua Wang [email protected] National Laboratory of Solid State Microstructures &\& School of Physics, Nanjing University, Nanjing 210093, China Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China
Abstract

There is a hot debate on the anomalous behavior of superfluid density ρs\rho_{s} in overdoped La2-xSrxCuO4 films in recent years. The linear drop of ρs\rho_{s} at low temperatures implies the superconductors are clean, but the linear scaling between ρs\rho_{s} (in the zero temperature limit) and the transition temperature TcT_{c} is a hallmark of the dirty limit in the Bardeen-Cooper-Schrieffer (BCS) framework [I. Bozovic et al., Nature 536, 309 (2016)]. This dichotomy motivated exotic theories beyond the standard BCS theory. We show, however, that such a dichotomy can be reconciled naturally by the role of increasing anisotropic scattering caused by the apical oxygen vacancies. Furthermore, the anisotropic scattering also explains the “missing” Drude weight upon doping in the optical conductivity, as reported in the THz experiment [F. Mahmood et al., Phys. Rev. Lett. 122, 027003 (2019)]. Therefore, the overdoped cuprates can actually be described consistently by the dd-wave BCS theory with the unique anisotropic scattering.

Introduction.— For a long time, the overdoped cuprate superconductors were believed to be described quite well by the Bardeen-Cooper-Schrieffer (BCS) theory Lee et al. (2006); Keimer et al. (2015). However, in 2016, such a belief was challenged by the measurement of the superfluid density ρs\rho_{s} using mutual inductance technique on a large amount of high quality overdoped La2-xSrxCuO4 (LSCO) films Božović et al. (2016). Two seemingly contradicting results were reported: ρs(0)ρs(T)T\rho_{s}(0)-\rho_{s}(T)\propto T and ρs(0)Tc\rho_{s}(0)\propto T_{c} where ρs(0)\rho_{s}(0) is the zero temperature value of ρs(T)\rho_{s}(T) and TcT_{c} is the transition temperature. Within the conventional BCS theory, the former scaling law indicates the d-wave superconductors are very clean, but the latter is a result of dirty BCS superconductors Abrikosov et al. (1963). This dichotomy regarding the dirtiness/cleanness of the overdoped cuprates motivated exotic theoretical investigations Božović et al. (2019); Zaanen (2016); Herman and Hlubina (2018); Fu et al. (2018); Wang (2018); Goutéraux and Mefford (2020); Phillips et al. (2020); Li et al. (2021).

The superfluid density has also been measured by THz optical conductivity experiment Mahmood et al. (2019), and is quantitatively consistent with the mutual inductance measurement Božović et al. (2016). In the meantime, the Drude fitting of the optical conductivity σ1(ν)\sigma_{1}(\nu) indicates the dirty limit Lee-Hone et al. (2017, 2018). So the same dichotomy also exists in optical measurements. Moreover, as yet another puzzle, σ1(ν0)\sigma_{1}(\nu\to 0) should be identical to, but is fitted to be significantly smaller than the dc conductivity σdc\sigma_{dc} Mahmood et al. (2019). This superficial loss of Drude weight seems to increase with overdoping.

In this Letter, we propose a scenario to resolve all of the above mysteries. We realize that in addition to the conventional isotropic scattering rate Γs\Gamma_{s} Lee-Hone et al. (2017, 2018), the apical oxygen vacancies cause an anisotropic scattering rate Γdcos2(2θ)\Gamma_{d}\cos^{2}(2\theta), with θ\theta the azimuthal angle of the quasiparticle momentum relative to the antinodal direction. Since the low-energy nodal quasiparticles are barely affected by Γd\Gamma_{d}, they reduce the superfluid density linearly in temperature if in addition Γs0\Gamma_{s}\rightarrow 0. But the total scattering rate Γθ=Γs+Γdcos2(2θ)\Gamma_{\theta}=\Gamma_{s}+\Gamma_{d}\cos^{2}(2\theta) determines the typical behavior ρs(0)Tc\rho_{s}(0)\propto T_{c} in the dirty limit. Meanwhile, the strong anisotropy in Γθ\Gamma_{\theta} causes a continuous distribution of Lorentzian components in σ(ν)\sigma(\nu), so that σ1(ν0)\sigma_{1}(\nu\to 0) would be underestimated by extrapolation from finite frequencies if a single isotropic scattering rate were assumed instead Mahmood et al. (2019).

Refer to caption
Figure 1: Illustration of apical oxygen coupling to (effective) dx2y2d_{x^{2}-y^{2}}-ortibals in the CuO2 plane. Hoppings with opposite signs are labeled by blue and red colors, respectively.

Oxygen vacancies and “cold spot” model.— From both early Torrance et al. (1988); Sato et al. (2000) and recent Kim et al. (2017) studies, apical oxygen vacancies were found to be an important factor to prevent obtaining (high-quality) overdoped LSCO samples. Therefore, the effect of apical oxygen vacancies deserves study in overdoped cuprates carefully. This is another motivation of the present work.

In Fig. 1, we plot a configuration of one apical oxygen atom above the CuO2-plane (only Cu-sites plotted). Notice that since the carriers have dx2y2d_{x^{2}-y^{2}}-like orbital content Zhang and Rice (1988), there is no coupling between the oxygen and the dx2y2d_{x^{2}-y^{2}}-orbital just below it. Therefore, the leading couplings are with the next nearest neighboring sites, giving the hopping integrals switching signs alternatively, as shown in Fig. 1. First, if there is no oxygen vacancy, the second order process of these out-of-plane hoppings gives an additional band dispersion 4κ(coskxcosky)2-4\kappa(\cos k_{x}-\cos k_{y})^{2} where κ=to2/εo\kappa=t_{o}^{2}/\varepsilon_{o} with tot_{o} and εo\varepsilon_{o} the hopping amplitude and energy level distance between the apical oxygen and dx2y2d_{x^{2}-y^{2}} orbitals Xiang and Wheatley (1996). Such a dispersion has already enters the carrier band. Next, we consider one oxygen vacancy at the origin. Relative to the uniform background, the vacancy now leads to an additional term Hi=δδσκfδfδcδσcδσH_{i}=\sum_{\delta\delta^{\prime}\sigma}\kappa f_{\delta}f_{\delta^{\prime}}c_{\delta\sigma}^{\dagger}c_{\delta^{\prime}\sigma} where fδ=1(1)f_{\delta}=1(-1) for δ=±x^(±y^)\delta=\pm\hat{x}(\pm\hat{y}) and σ\sigma denotes spin, which gives rise to the following scattering Hamiltonian

Hi=kkσ4κN(coskxcosky)(coskxcosky)ckσckσ\displaystyle H_{i}=\sum_{kk^{\prime}\sigma}\frac{4\kappa}{N}(\cos k_{x}-\cos k_{y})(\cos k_{x}^{\prime}-\cos k_{y}^{\prime})c_{k\sigma}^{\dagger}c_{k^{\prime}\sigma} (1)

where kk and kk^{\prime} are momentums, and NN is the number of copper atoms. Since the oxygen vacancies are out-of-plane and only couple to next nearest neighboring dx2y2d_{x^{2}-y^{2}}-orbitals, the scattering strength is anticipated to be small. Under the Born approximation Abrikosov et al. (1963), HiH_{i} gives a scattering rate Γ(kx,ky)Γd(coskxcosky)2/4\Gamma(k_{x},k_{y})\propto\Gamma_{d}(\cos k_{x}-\cos k_{y})^{2}/4. For analytical convenience but without loss of qualitative physics, throughout this work, we further assume circular fermi surface and use wide band approximation, so that the scattering rate is only angle-dependent, i.e.

Γθ=Γs+Γdcos2(2θ)\displaystyle\Gamma_{\theta}=\Gamma_{s}+\Gamma_{d}\cos^{2}(2\theta) (2)

where we have added the isotropic scattering rate Γs\Gamma_{s} arising from, e.g., in-plane impurities. In the following calculations, Γs\Gamma_{s} and Γd\Gamma_{d} are our model parameters. In fact, this kind of scattering rate has been proposed to explain the transport phenomena in underdoped cuprates, called “cold spot” model Ioffe and Millis (1998), where the Γd\Gamma_{d}-term is attributed to the pair fluctuations or interaction effects, which hence is expected to decrease upon doing. But here, in our case, the Γd\Gamma_{d}-term is caused by oxygen vacancies and should increase upon doping, according to both early Torrance et al. (1988); Sato et al. (2000) and recent Kim et al. (2017) experiments.

Refer to caption
Figure 2: (a) The DOS ρ\rho versus ω\omega in the case of pure isotropic scattering Γs=0.2Δ0\Gamma_{s}=0.2\Delta_{0} (solid line). For comparison, the dashed line shows the DOS in the clean limit. (b) The same as (a) but in the case of pure anisotropic scattering Γd=0.2Δ0\Gamma_{d}=0.2\Delta_{0}. The shaded region highlights the similarity to the clean limit. (c) and (d) are angle-dependent PDOS in the presence of scattering corresponding to (a) and (b), respectively. (e) Tc/Tc0T_{c}/T_{c0} versus Γs/Tc0\Gamma_{s}/T_{c0} (blue) and Γd/Tc0\Gamma_{d}/T_{c0} (red) for the two types of scatterings. (f) The ratio 2Δ0(0)/Tc2\Delta_{0}(0)/T_{c} versus Tc/Tc0T_{c}/T_{c0} (red and blue dots) for the two cases considered in (e), in comparison to the clean limit result (dashed line).

We now look at the effect of the anisotropic scattering Γd\Gamma_{d} on single particle excitations. By choosing the d-wave pairing function Δ(θ)=Δ0cos(2θ)\Delta(\theta)=\Delta_{0}\cos(2\theta) and fixing Γs,d=0.2Δ0\Gamma_{s,d}=0.2\Delta_{0}, we calculate the density of states (DOS) ρ(ω)=dθ2πρ(ω,θ)\rho(\omega)=\int\frac{\mathrm{d}\theta}{2\pi}\rho(\omega,\theta), where ρ(ω,θ)\rho(\omega,\theta) is the partial DOS (PDOS) given by

ρ(ω,θ)=𝒩πdε[Γθ(ωEθ)2+Γθ2+Γθ(ω+Eθ)2+Γθ2]\displaystyle\rho(\omega,\theta)=\frac{\mathcal{N}}{\pi}\int\mathrm{d}\varepsilon\left[\frac{\Gamma_{\theta}}{(\omega-E_{\theta})^{2}+\Gamma_{\theta}^{2}}+\frac{\Gamma_{\theta}}{(\omega+E_{\theta})^{2}+\Gamma_{\theta}^{2}}\right] (3)

where 𝒩\mathcal{N} is the normal state DOS and Eθ=ε2+Δ(θ)2E_{\theta}=\sqrt{\varepsilon^{2}+\Delta(\theta)^{2}}. ρ(ω)\rho(\omega) and ρ(ω,θ)\rho(\omega,\theta) are plotted in Fig. 2(a)-(d). In the calculations, we perform the integrals over ε\varepsilon and θ\theta using standard numerical integral technique. Fig. 2(a) is the typical behavior of the conventional d-wave superconductor with isotropic scattering rate: zero energy cusp and coherence peak are both smeared out by isotropic scattering rate. Differently, the Γd\Gamma_{d}-scattering shown in Fig. 2(b) only suppresses the coherence peak but does not change the zero energy cusp, as a result of the vanishing of scattering rate for nodal quasiparticles. Accordingly, we can divide the energy space into two regions: lower energy “clean” one and higher energy “dirty” one. The above picture is also seen clearly from the PDOS in Fig. 2(d), where the spectral becomes smeared out for high energy quasiparticles (near the antinodes) but remains sharp for low energy ones (near the nodes).

The dirty BCS theory for d-wave superconductors Alloul et al. (2009) is a direct generalization of the Abrikosov-Gor’kov (AG) theory Abrikosov and Gor’kov (1961), since the potential scattering here causes pair-breaking which is similar to the magnetic impurities in s-wave superconductors. The gap Δ0\Delta_{0} is determined by the self-consistent equation

1=λTωn>0Ωdθϕθ2ωn2+2|ωn|Γθ+Γθ2+Δ02ϕθ2\displaystyle 1=\lambda T\sum_{\omega_{n}>0}^{\Omega}\int\mathrm{d}\theta\frac{\phi_{\theta}^{2}}{\sqrt{\omega_{n}^{2}+2|\omega_{n}|\Gamma_{\theta}+\Gamma^{2}_{\theta}+\Delta_{0}^{2}\phi_{\theta}^{2}}} (4)

where λ\lambda is the BCS coupling constant, ωn=(2n1)πT\omega_{n}=(2n-1)\pi T the Matsubara frequency, and ϕθ=cos2θ\phi_{\theta}=\cos 2\theta the d-wave form factor. 111Exactly speaking, the frequency summation should be bounded with a soft-cutoff (with the form of a boson propagator) instead of the hard-cutoff used here. However, only the latter can be mapped to the momentum cutoff strategy as in the BCS treatment. In numerical calculations, we use λ=0.3\lambda=0.3 and Ω=800\Omega=800, which gives Tc0=1.134Ωe2/λ=1.15T_{c0}=1.134\Omega\mathrm{e}^{-2/\lambda}=1.15 for clean superconductors (with this setup, 0.87Tc00.87T_{c0} is taken as the energy unit). By letting Δ00\Delta_{0}\to 0, we obtain the generalized TcT_{c}-formula

lnTc0Tc=1πdθϕθ2[ψ(12+Γθ2πTc)ψ(12)]\displaystyle\ln\frac{T_{c0}}{T_{c}}=\frac{1}{\pi}\int\mathrm{d}\theta\phi_{\theta}^{2}\left[\psi\left(\frac{1}{2}+\frac{\Gamma_{\theta}}{2\pi T_{c}}\right)-\psi\left(\frac{1}{2}\right)\right] (5)

where ψ(z)\psi(z) is the digamma function. The results of TcT_{c} vs Γs\Gamma_{s} and Γd\Gamma_{d} are shown in Fig. 2(e). A stronger Γd\Gamma_{d} is needed to kill superconductivity as the low energy excitations are less affected. In Fig. 2(f), we show the results of 2Δ0/Tc2\Delta_{0}/T_{c}. It is interesting to find that the anisotropic scattering drives the ratio farther away from the BCS prediction 4.284.28 in the clean limit to values as large as 10\sim 10. In the literature, a large gap-TcT_{c} ratio is often taken as an indication of strong coupling superconductivity. Here, the anisotropic scattering gives another interpretation.

Superfluid density.— After recognizing the clean-dirty dichotomy caused by oxygen vacancies at the single particle level, we turn to discuss the superfluid density, which can be obtained as Coleman (2015)

ρs=e2𝒩vF2TωndθΔθ2cos2θ(ωn2+2|ωn|Γθ+Γθ2+Δθ2)3/2\displaystyle\rho_{s}=e^{2}\mathcal{N}v_{F}^{2}T\sum_{\omega_{n}}\int\mathrm{d}\theta\frac{\Delta_{\theta}^{2}\cos^{2}\theta}{\left(\omega_{n}^{2}+2|\omega_{n}|\Gamma_{\theta}+\Gamma_{\theta}^{2}+\Delta_{\theta}^{2}\right)^{3/2}} (6)

where vFv_{F} is the Fermi velocity. Here, the current vertex correction can be proven to vanish in the non-crossing approximation as a result of the factorized scattering potential (Eq. 1), as shown in the Supplementary Material sm . In Fig. 6(a) and (b), ρs\rho_{s} vs TT is plotted for pure Γs\Gamma_{s} and Γd\Gamma_{d} scatterings, respectively. The most obvious difference is that any nonzero Γs\Gamma_{s} causes power law temperature dependence of ρs(T)\rho_{s}(T), but Γd\Gamma_{d} preserves the linear dependence as in the clean limit. This is already anticipated from the DOS feature in Fig. 2(b) because the normal fluid density ρn\rho_{n} is roughly determined by quasiparticles and thus satisfies ρn(T)ρn(0)T\rho_{n}(T)-\rho_{n}(0)\propto T, leaving the superfluid density satisfying ρs(0)ρs(T)=ρn(T)ρn(0)T\rho_{s}(0)-\rho_{s}(T)=\rho_{n}(T)-\rho_{n}(0)\propto T.

Refer to caption
Figure 3: (a) Superfluid density ρs/ρs0\rho_{s}/\rho_{s0} as a function of T/Tc0T/T_{c0} in the case of isotropic scattering (Γd=0\Gamma_{d}=0), for various values of Γs\Gamma_{s} which increases uniformly from top to bottom. Here ρs0=ρs(0)\rho_{s0}=\rho_{s}(0) in the clean limit. (b) The same as (a) but in the case of anisotropic scattering (Γs=0\Gamma_{s}=0) and for various values of Γd\Gamma_{d}. (c) ρs/ρs(0)\rho_{s}/\rho_{s}(0) as a function of T/TcT/T_{c} for (Γs,Γd)=(0,0)(\Gamma_{s},\Gamma_{d})=(0,0) (dashed blue), (0.1,0)(0.1,0) (green), (1,0)(1,0) (dashed pink), (0,1.2)(0,1.2) (blue). The red dots show all 100100 groups of experimental data Božović et al. (2016) for comparison. The inset shows the fit to the experimental data with Tc=5.1T_{c}=5.1K using (Γs,Γd)=(0,0.75)(\Gamma_{s},\Gamma_{d})=(0,0.75). (d) ρs(0)/ρs0\rho_{s}(0)/\rho_{s0} as a function of Tc/Tc0T_{c}/T_{c0}. The lines are obtained by fixing one of (Γs,Γd)(\Gamma_{s},\Gamma_{d}) (as indicated) while varying the other. The inset shows two similar curves by fixing Γs=0\Gamma_{s}=0 and Γd=0\Gamma_{d}=0, respectively, in a broader scale. The results of all other cases (not shown) fall within the shaded regime bounded by these two curves. In (c) and (d), the insets share the same axis labels as for the main panels.

In real materials, both Γs\Gamma_{s} and Γd\Gamma_{d} are expected to coexist. In order to make quantitative comparisons with the experiment, we renormalize TT and ρs\rho_{s} by TcT_{c} and ρs(0)\rho_{s}(0), respectively, as shown in Fig. 6(c). We have superposed the experimental data of all 100100 samples with TcT_{c} ranging from 41.641.6 to 5.15.1K Božović et al. (2016), shown as the red dots. Four typical theoretical curves are plotted: (Γs,Γd)=(0,0)(\Gamma_{s},\Gamma_{d})=(0,0), (0.1,0)(0.1,0), (1,0)(1,0), and (0,1.2)(0,1.2). As can be seen, almost all the experimental data reside within the region enclosed by the curves of (0.1,0)(0.1,0) (green line) and (0,1.2)(0,1.2) (blue line), while if Γs\Gamma_{s} is large and dominant, as the (1,0)(1,0)-line shows, the renormalized ρs\rho_{s}-TT curve bends more significantly, which is at odds with the experimental data. These observations indicate the more important role of Γd\Gamma_{d}. As an example, the experimental data (open circles) with Tc=5.1T_{c}=5.1K are shown in the inset, which is clearly very linear. We can fit the data with pure Γd=0.75\Gamma_{d}=0.75 (line) quite well.

Another key observation of the experiment Božović et al. (2016) is the linear relationship between ρs(0)\rho_{s}(0) vs TcT_{c}. To see that, we plot our results in Fig. 6(d). Pure Γs\Gamma_{s} and Γd\Gamma_{d} scatterings correspond to blue and red curves, which enclose the physical region (shaded region in the inset), in which ρs(0)\rho_{s}(0) is always roughly proportional to TcT_{c} although not exactly. Therefore, this is a hallmark of the dirty BCS superconductors regardless of isotropic (Γs\Gamma_{s}) or anisotropic (Γd\Gamma_{d}) scatterings. Furthermore, we have also fix Γs=0.05\Gamma_{s}=0.05 and 0.10.1 and tune Γd\Gamma_{d} to suppress TcT_{c}. Both of them show the bending behavior near Tc0T_{c}\to 0, showing much better behavior than the separate scattering cases in view of the experimental data .

Refer to caption
Figure 4: (a) Optical conductivity σ1n\sigma_{1n} versus frequency ν\nu in the normal state for various cases of (Γs,Γd)(\Gamma_{s},\Gamma_{d}) (as indicated) in unit of Δ0\Delta_{0}. The inset is a replot of σ1\sigma_{1} in the main panel but normalized by σ1n(0)\sigma_{1n}(0). (b) The same as (a) but for the superconducting state conductivity σ1\sigma_{1}. The inset shows the result for isotropic scattering, revealing the universal conductivity in the zero frequency limit. (c) σ1/σ1n\sigma_{1}/\sigma_{1n} versus ν\nu, with σ1n\sigma_{1n} and σ1\sigma_{1} from (a) and (b), respectively. The inset is for isotropic scattering. In (a)-(c), the line color is associated with the scattering parameter setting identically, and in (b) and (c), the insets share the same axis labels as for the main panels. (d) Best fits (lines) to the experimental data (symbols) with Tc=17.5T_{c}=17.5K measured at 2222K (blue) and 1.61.6K (red) Mahmood et al. (2019). The dashed lines are best fits with Γs\Gamma_{s} alone, and the solid lines include both Γs\Gamma_{s} and Γd\Gamma_{d}.

Optical conductivity.— From the Kubo formula, the optical conductivity σ1(ν)\sigma_{1}(\nu) can be obtained as Hirschfeld et al. (1989)

σ1(ν)=\displaystyle\sigma_{1}(\nu)= 𝒩e2vF22π3dεdθdωf(ω)f(ω+ν)νcos2θ\displaystyle\frac{\mathcal{N}e^{2}v_{F}^{2}}{2\pi^{3}}\int\mathrm{d}\varepsilon\mathrm{d}\theta\mathrm{d}\omega\frac{f(\omega)-f(\omega+\nu)}{\nu}\cos^{2}\theta{}
tr[ImG(ω,ε,θ)ImG(ω+ν,ε,θ)]\displaystyle\text{tr}\left[\text{Im}G(\omega,\varepsilon,\theta)\text{Im}G(\omega+\nu,\varepsilon,\theta)\right] (7)

where f(ω)f(\omega) is the Fermi distribution function, G(ω,ε,θ)G(\omega,\varepsilon,\theta) is the Green’s function in the Nambu space: G1=(ω+iΓθ)σ0εσ3Δ0ϕθσ1G^{-1}=(\omega+i\Gamma_{\theta})\sigma_{0}-\varepsilon\sigma_{3}-\Delta_{0}\phi_{\theta}\sigma_{1} with σ0,1,3\sigma_{0,1,3} the identity and Pauli matrices. By numerical calculations, we obtain the frequency dependence of σ1(ν)\sigma_{1}(\nu) in both normal and superconducting states (with Δ0=1\Delta_{0}=1 as the energy unit), as shown in Fig. 7(a) and (b), respectively. In the normal state, as Γd\Gamma_{d} increases, σ1(ν)\sigma_{1}(\nu) becomes more and more broadened to show long tail behavior. This is an important feature obtained in the “cold spot” model Ioffe and Millis (1998). In the superconducting state, a nonzero Γd\Gamma_{d} breaks the universal conductivity Lee (1993), which applies only in the case of pure Γs\Gamma_{s} scattering as shown in the inset of Fig. 7(b). Furthermore, we renormalize the superconducting state σ1(ν)\sigma_{1}(\nu) by the normal state value σ1n(ν)\sigma_{1n}(\nu) to obtain the results shown in Fig. 7(c). The resulted curves show similar behavior to the DOS at low frequencies: nearly linear ν\nu-dependence if Γd\Gamma_{d} dominates which is quite different from the case of pure Γs\Gamma_{s} scattering (inset) with quadratic ν\nu-dependence. This can be used to examine the types of scattering in future experiments.

The most obvious feature when Γd\Gamma_{d} dominates is the behavior of σ1(ν)\sigma_{1}(\nu) is no longer a simple Lorentian-like Drude peak (as in the case of isotropic scattering), since the quasiparticles experience angle dependent scattering rates. In fact, the conductivity should be an integration over a continuous distribution of Lorentzian components. Since the Lorentzian is sharper and higher for smaller scattering rates, the overall line shape of σ1(ν)\sigma_{1}(\nu) should be steeper as the frequency approaches zero. Therefore, if the finite frequency data are used to perform the fit to a single Lorentzian-like Drude peak, as in the THz experiment, the extrapolated value of σ1(ν0)\sigma_{1}(\nu\to 0) will be lower than the actual one, leading to superficial “missing” of the Drude weight. As an example, in Fig. 7(d) we fit the experimental data with Tc=17.5T_{c}=17.5K (symbols) in both normal (blue) and superconducting (red) states Mahmood et al. (2019). The best fits with only Γs\Gamma_{s} (dashed lines) are found to underestimate σ1(ν0)\sigma_{1}(\nu\to 0) and the spectral weight, as compared to the result with both Γs\Gamma_{s} and Γd\Gamma_{d} (solid lines). Since the pairing gap Δ0\Delta_{0} is not available in the experiment, we take it also as a parameter, and the fitted value is 1.111.11THz. More systematic fitting for a series of overdoped samples can be found in the Supplemental Material sm , which supports our main point that apical oxygen vacancy increases with overdoping, leading to larger Γd\Gamma_{d} scattering.

Summary and discussions.— In summary, we have found that the apical oxygen vacancies give rise to an anisotropic scattering rate Γdcos2(2θ)\Gamma_{d}\cos^{2}(2\theta) which causes a clean-dirty dichotomy for low-high energy excitations. This provides a natural explanation of the anomalous behavior of the superfluid density and THz optical conductivity in the experiments. Therefore, we conclude that the superconducting states of overdoped cuprates are still captured by the BCS theory, as long as the anisotropic scattering rate is adequately considered.

Finally, we make some remarks. (1) In this work, we have omitted many material-dependent details (such as specific tight-binding models, pairing interactions and interaction effects) in order to obtain universal results. The doping dependence enters the problem via the scattering rate Γd\Gamma_{d}, which increases with doping. (2) The anisotropic scattering rate has been reported in angle-dependent magnetoresistivity experiments in Nd-LSCO Grissonnanche et al. (2021) and Tl2Ba2CuO6+δ Abdel-Jawad et al. (2006). But to the best of our knowledge, there has not been such report in overdoped LSCO films. A systematic study of the anisotropic scattering rate can check our model and finally answer the question on whether overdoped cuprates are dirty BCS superconductors or not. (3) Our prediction suggests that it is necessary to extend the optical measurements down to even lower frequency (e.g. GHz) in order to obtain accurate behavior of the optical conductivity. (4) About superfluid density, there are two other kinds of scalings reported in literature: Uemura’s law Uemura et al. (1989) ρs(0)Tc\rho_{s}(0)\propto T_{c} and Homes’ law ρs(0)σdcTc\rho_{s}(0)\propto\sigma_{dc}T_{c} Homes et al. (2004). The former is obtained in underdoped cuprates, where phase fluctuations are very strong and cannot be neglected. Emery and Kivelson (1995). On the other hand, Homes’ law can be explained by dirty BCS superconductors with pair-preserving impurities only Abrikosov et al. (1963); Kogan (2013). Here, in the overdoped LSCO films, the oxygen vacancies are pair-breaking, hence Homes’ law does not apply. (5) Recently, the apical oxygen vacancies have also been observed in optimally doped YBa2Cu3O7-x Hartman et al. (2019). Then, according to our study, σ1(ν)\sigma_{1}(\nu) at low frequency should be non-standard Drude-like, which may be consistent with the early microwave experiment Turner et al. (2003). Furthermore, if the apical oxygen vacancies are widely present in different families of cuprates, then the effect of Γd\Gamma_{d} scattering should be considered carefully in future studies.

D. W. thanks Congjun Wu, Qi Zhang, Tao Li, Yuan Wan for helpful discussions during the past few years. D. W. also acknowledges Ivan Bozovic for his encouragement on studying their experiments. This work is supported by National Natural Science Foundation of China (under Grant Nos. 11874205, 11574134, and 12074181), National Key Projects for Research and Development of China (Grant No.2021YFA1400400), Fundamental Research Funds for the Central Universities (Grant No. 020414380185), and Natural Science Foundation of Jiangsu Province (No. BK20200007).

References

Supplementary Materials

In this supplementary material, we first derive the scattering potential caused by the apical oxygen vacancies and then the anisotropic scattering rate Γdcos2(2θ)\Gamma_{d}\cos^{2}(2\theta) within the Born approximation. The effects of this scattering rate on some superconducting properties, including density of states, gap, TcT_{c}, superfluid density, and optical conductivity, are derived in a self-contained way. Finally, some further discussions are given, including first principle calculations to determine the position of the oxygen vacancies (apical or planar), the effect of planar oxygen vacancies (if they exist), and the current vertex corrections.

I Disorder model of apical oxygen vacancies

I.1 Scattering potential

At first, suppose there is no oxygen vacancy, we have the following translational invariant Hamiltonian

Hdp=\displaystyle H_{dp}= ij(tijdidj+h.c.)ito(pidi+x^+pidix^pidi+y^pidiy^+h.c.)\displaystyle-\sum_{\left\langle ij\right\rangle}(t_{ij}d_{i}^{\dagger}d_{j}+h.c.)-\sum_{i}t_{o}(p_{i}^{\dagger}d_{i+\hat{x}}+p_{i}^{\dagger}d_{i-\hat{x}}-p_{i}^{\dagger}d_{i+\hat{y}}-p_{i}^{\dagger}d_{i-\hat{y}}+h.c.){}
+i(εddidi+εopipi)\displaystyle+\sum_{i}(\varepsilon_{d}d_{i}^{\dagger}d_{i}+\varepsilon_{o}p_{i}^{\dagger}p_{i}) (8)

where did_{i}^{\dagger} generates the in-plane electrons effectively at the Cu-site and pip_{i}^{\dagger} generates the ones at the apical O-site. The sign change of the out-of-plane hopping (with amplitude tot_{o}) between ±x^\pm\hat{x} and ±y^\pm\hat{y} are caused by the (x2y2)(x^{2}-y^{2}) orbital character of the in-plane carriers (either dx2y2d_{x^{2}-y^{2}} orbital electrons or Zhang-Rice singlets), see Fig. 1 in the main text for clarity. We have omitted the spin indices here for simplicity because they are irrelevant to obtain the scattering potential. After integrating out the pp-electrons (or by a second order perturbation), we obtain

Hd=ij(tijdidj+h.c.)κiδ=±x^,±y^fδfδdi+δdi+δ+iεddidi\displaystyle H_{d}=-\sum_{\left\langle ij\right\rangle}(t_{ij}d_{i}^{\dagger}d_{j}+h.c.)-\kappa\sum_{i}\sum_{\delta=\pm\hat{x},\pm\hat{y}}f_{\delta}f_{\delta^{\prime}}d_{i+\delta}^{\dagger}d_{i+\delta^{\prime}}+\sum_{i}\varepsilon_{d}d_{i}^{\dagger}d_{i} (9)

where κ=to2εo\kappa=\frac{t_{o}^{2}}{\varepsilon_{o}}, fδ=1f_{\delta}=1 for δ=±x^\delta=\pm\hat{x} and fδ=1f_{\delta}=-1 for δ=±y^\delta=\pm\hat{y}. The Fourier transformation then gives the band dispersion

εk=εkd4κ(coskxcosky)2\displaystyle\varepsilon_{k}=\varepsilon_{kd}-4\kappa(\cos k_{x}-\cos k_{y})^{2} (10)

where the first term εkd\varepsilon_{kd} is from the pure dd-electrons, and the second term comes from the coupling to apical oxygen orbitals. In together, εk\varepsilon_{k} is taken as the free band without any disorder.

Next, let us consider one apical oxygen vacancy at the origin. Relative to the translational invariant background, the single vacancy is expressed by the impurity Hamiltonian

Himp=κδ=±x^,±y^fδfδdδdδ\displaystyle H_{\rm imp}=\kappa\sum_{\delta=\pm\hat{x},\pm\hat{y}}f_{\delta}f_{\delta^{\prime}}d_{\delta}^{\dagger}d_{\delta^{\prime}} (11)

whose Fourier transformation then gives a scattering potential

Himp=4κNkk(coskxcosky)(coskxcosky)dkdk\displaystyle H_{\rm imp}=\frac{4\kappa}{N}\sum_{kk^{\prime}}(\cos k_{x}-\cos k_{y})(\cos k_{x}^{\prime}-\cos k_{y}^{\prime})d_{k}^{\dagger}d_{k^{\prime}} (12)

which is Eq. 1 in the main text.

I.2 Anisotropic scattering rate: Born approximation

As explained in the main text, since the impurity is out-of-plane, its intensity is expected to be quite small such that the Born approximation is acceptable.

{feynman}\vertex\vertex\vertex\vertex\vertexnimpn_{\rm imp}\diagramkiωnk\,i\omega_{n}kiωnk^{\prime}\,i\omega_{n}kiωnk\,i\omega_{n}VkkV_{kk^{\prime}}VkkV_{k^{\prime}k}
Figure 5: The single particle self energy in the Born approximation.

The self energy under the Born approximation Abrikosov et al. (1963) as shown in Fig. 5 is

Σ(k,iωn)=nimpd2k(2π)2|Vkk|2G(k,iωn)\displaystyle\Sigma(k,i\omega_{n})=n_{\rm imp}\int\frac{\mathrm{d}^{2}k}{(2\pi)^{2}}|V_{kk^{\prime}}|^{2}G(k^{\prime},i\omega_{n}) (13)

where G(k,iωn)=(iωnεk)1G(k,i\omega_{n})=(i\omega_{n}-\varepsilon_{k})^{-1}. As usual, for simplicity, we adopt the wide band approximation with circular Fermi surface and keep the form factor (coskxcosky)(\cos k_{x}-\cos k_{y}) with only the angle dependence 2cos(2θ)2\cos(2\theta), we obtain

Σ(θ,iωn)\displaystyle\Sigma(\theta,i\omega_{n}) =256κ2nimp𝒩cos2(2θ)dεdθ2πcos2(2θ)iωnε\displaystyle=256\kappa^{2}n_{\rm imp}\mathcal{N}\cos^{2}(2\theta)\int\mathrm{d}\varepsilon\int\frac{\mathrm{d}\theta^{\prime}}{2\pi}\frac{\cos^{2}(2\theta^{\prime})}{i\omega_{n}-\varepsilon}{}
=i128πκ2nimp𝒩sgn(ωn)cos2(2θ)\displaystyle=-i128\pi\kappa^{2}n_{\rm imp}\mathcal{N}\text{sgn}(\omega_{n})\cos^{2}(2\theta) (14)

by analytic continuation, it gives the desired scattering rate Γθ=Γdcos2(2θ)\Gamma_{\theta}=\Gamma_{d}\cos^{2}(2\theta) with Γd=128πκ2nimp𝒩\Gamma_{d}=128\pi\kappa^{2}n_{\rm imp}\mathcal{N}. Together with the usual isotropic scattering rate Γs\Gamma_{s}, we get Eq. 2 in the main text.

II Disorder effect

II.1 Density of states

The retarded Green’s function in the superconducting state can be written in the Nambu space as GR1(ω,k)=(ω+iΓk)σ0εkσ3Δkσ1G_{R}^{-1}(\omega,k)=(\omega+i\Gamma_{k})\sigma_{0}-\varepsilon_{k}\sigma_{3}-\Delta_{k}\sigma_{1}, where σ0,1,3\sigma_{0,1,3} are identity and Pauli matrices. Keeping only the angle dependence and after integration over ε\varepsilon, we obtain

GR(ω,θ)\displaystyle G_{R}(\omega,\theta) =𝒩dεGR(ω,θ,ε)\displaystyle=\mathcal{N}\int\mathrm{d}\varepsilon G_{R}(\omega,\theta,\varepsilon){}
=𝒩2dε(1ω+iΓθEθ+1ω+iΓθ+Eθ)σ0\displaystyle=\frac{\mathcal{N}}{2}\int\mathrm{d}\varepsilon\left(\frac{1}{\omega+i\Gamma_{\theta}-E_{\theta}}+\frac{1}{\omega+i\Gamma_{\theta}+E_{\theta}}\right)\sigma_{0} (15)

The angle-dependent partial density of states (PDOS) ρ(ω,θ)\rho(\omega,\theta) is defined as the imaginary part of GR(ω,θ)G_{R}(\omega,\theta), hence, given by

ρ(ω,θ)=2πIm[GR(ω,θ)]11=𝒩πdε[Γθ(ωEθ)2+Γθ2+Γθ(ω+Eθ)2+Γθ2]\displaystyle\rho(\omega,\theta)=-\frac{2}{\pi}\text{Im}[G_{R}(\omega,\theta)]_{11}=\frac{\mathcal{N}}{\pi}\int\mathrm{d}\varepsilon\left[\frac{\Gamma_{\theta}}{(\omega-E_{\theta})^{2}+\Gamma_{\theta}^{2}}+\frac{\Gamma_{\theta}}{(\omega+E_{\theta})^{2}+\Gamma_{\theta}^{2}}\right] (16)

which is Eq. 3 in the main text. The angle integral of ρ(ω,θ)\rho(\omega,\theta) then gives the total density of states (DOS) ρ(ω)=dθ2πρ(ω,θ)\rho(\omega)=\int\frac{\mathrm{d}\theta}{2\pi}\rho(\omega,\theta).

It is interesting to notice that near the nodal region θ=π/4δθ\theta=\pi/4-\delta\theta, the d-wave pairing Δθ=Δ0cos(2θ)Δ0δθ\Delta_{\theta}=\Delta_{0}\cos(2\theta)\approx\Delta_{0}\delta\theta vanishes linearly with δθ\delta\theta, while the Γd\Gamma_{d}-scattering vanishes quadratically ΓθΓdδθ2\Gamma_{\theta}\approx\Gamma_{d}\delta\theta^{2}. Therefore, the Γd\Gamma_{d}-scattering has little effect on the low energy quasiparticles, which is quite different from the isotropic scattering and causes the clean-dirty dichotomy as we discussed in the main text.

II.2 Gap and TcT_{c}

Within the BCS framework, we assume a pairing interaction

HI=Vkkϕkϕkdkdkdkdk\displaystyle H_{I}=-V\sum_{kk^{\prime}}\phi_{k}\phi_{k^{\prime}}d_{k\uparrow}^{\dagger}d_{-k\downarrow}^{\dagger}d_{-k^{\prime}\downarrow}d_{k^{\prime}\uparrow} (17)

where ϕk=(coskxcosky)/2\phi_{k}=(\cos k_{x}-\cos k_{y})/2, correspondingly ϕθ=cos(2θ)\phi_{\theta}=\cos(2\theta). Then, the gap function Δk=Δ0ϕk\Delta_{k}=\Delta_{0}\phi_{k} is determined by the self-consistent condition,

Δk\displaystyle\Delta_{k^{\prime}} =VTϕkNωn,kϕkG21(ωn,k)\displaystyle=-\frac{VT\phi_{k^{\prime}}}{N}\sum_{\omega_{n},k}\phi_{k}G_{21}(\omega_{n},k){}
=2VTϕkNωn>0,kΔ0ϕk2(ωn+Γk)2+εk2+Δ02ϕk2\displaystyle=\frac{2VT\phi_{k^{\prime}}}{N}\sum_{\omega_{n}>0,k}\frac{\Delta_{0}\phi_{k}^{2}}{(\omega_{n}+\Gamma_{k})^{2}+\varepsilon_{k}^{2}+\Delta_{0}^{2}\phi_{k}^{2}}{}
=2πϕkV𝒩Tωn>0dθ2πΔ0ϕθ2(ωn+Γθ)2+Δ02ϕθ2\displaystyle=2\pi\phi_{k^{\prime}}V\mathcal{N}T\sum_{\omega_{n}>0}\int\frac{\mathrm{d}\theta}{2\pi}\frac{\Delta_{0}\phi_{\theta}^{2}}{\sqrt{(\omega_{n}+\Gamma_{\theta})^{2}+\Delta_{0}^{2}\phi_{\theta}^{2}}} (18)

which gives Eq. 4 in the main text by defining λ=V𝒩\lambda=V\mathcal{N}. In the last line, we have also adopted the wide band approximation with circular Fermi surface.

Next, by letting Δ00\Delta_{0}\to 0, we can determine TcT_{c}. The calculation is in parallel to Ref. Abrikosov and Gor’kov, 1961. The above self-consistent equation becomes

1\displaystyle 1 =2πλTcωn>0Ωdθ2πϕθ2ωn+Γθ\displaystyle=2\pi\lambda T_{c}\sum_{\omega_{n}>0}^{\Omega}\int\frac{\mathrm{d}\theta}{2\pi}\frac{\phi_{\theta}^{2}}{\omega_{n}+\Gamma_{\theta}}{}
=2πλTcωn>0Ωdθ2π(ϕθ2ωn+Γθϕθ2ωn+ϕθ2ωn)\displaystyle=2\pi\lambda T_{c}\sum_{\omega_{n}>0}^{\Omega}\int\frac{\mathrm{d}\theta}{2\pi}\left(\frac{\phi_{\theta}^{2}}{\omega_{n}+\Gamma_{\theta}}-\frac{\phi_{\theta}^{2}}{\omega_{n}}+\frac{\phi_{\theta}^{2}}{\omega_{n}}\right){}
=λn=0dθ2πϕθ2(1n+12+Γθ2πTc1n+12)+2πλTcωn>0Ωdθ2πϕθ2ωn\displaystyle=\lambda\sum_{n=0}^{\infty}\int\frac{\mathrm{d}\theta}{2\pi}\phi_{\theta}^{2}\left(\frac{1}{n+\frac{1}{2}+\frac{\Gamma_{\theta}}{2\pi T_{c}}}-\frac{1}{n+\frac{1}{2}}\right)+2\pi\lambda T_{c}\sum_{\omega_{n}>0}^{\Omega}\int\frac{\mathrm{d}\theta}{2\pi}\frac{\phi_{\theta}^{2}}{\omega_{n}} (19)

where we have added and subjected the same term (with Γθ=0\Gamma_{\theta}=0) such that the summation of the first two terms converges and upper limit can now be pushed to infinity. The last term is the same as the clean case without disorder except Tc0T_{c0} replaced by TcT_{c}. Using the clean case result Tc0=αΩe2/λT_{c0}=\alpha\Omega\mathrm{e}^{-2/\lambda} for d-wave pairing, we obtain

1λ2πTcωn>0Ωdθ2πϕθ2ωn=12lnTcTc0\displaystyle\frac{1}{\lambda}-2\pi T_{c}\sum_{\omega_{n}>0}^{\Omega}\int\frac{\mathrm{d}\theta}{2\pi}\frac{\phi_{\theta}^{2}}{\omega_{n}}=\frac{1}{2}\ln\frac{T_{c}}{T_{c0}} (20)

substituting it into Eq. 19, we obtain

lnTc0Tc=dθπϕθ2[ψ(12+Γθ2πTc)ψ(12)]\displaystyle\ln\frac{T_{c0}}{T_{c}}=\int\frac{\mathrm{d}\theta}{\pi}\phi_{\theta}^{2}\left[\psi\left(\frac{1}{2}+\frac{\Gamma_{\theta}}{2\pi T_{c}}\right)-\psi\left(\frac{1}{2}\right)\right] (21)

which is Eq. 5 in the main text.

II.3 Superfluid density

Next, let us consider the superfluid density. The calculation mainly follows Ref. Coleman, 2015. The superfluid density is given by two diagrams as shown in Fig. 6. The first diagram is from the paramagnetic current JkP=eεkdkdkJ_{k}^{P}=e\nabla\varepsilon_{k}d_{k}^{\dagger}d_{k} and the second one is from the diamagnetic current Jkd=e22εkdkdkJ_{k}^{d}=e^{2}\nabla^{2}\varepsilon_{k}d_{k}^{\dagger}d_{k}.

{feynman}\vertexvkσ0v_{k}\sigma_{0}vkσ0v_{k}\sigma_{0}\vertex\diagramkiωnk\,i\omega_{n}kiωnk\,i\omega_{n}
\feynmandiagram

[layered layout] a – [photon] b [dot] – [fermion,out=135, in=45, loop, min distance=4cm, edge label=kiωnk\,i\omega_{n}] b [label=270:2εkσ3\nabla^{2}\varepsilon_{k}\sigma_{3}] – [photon] c, ;

Figure 6: Feynman diagrams for the superfluid density ρs\rho_{s}.

Put these two contributions together, we obtain the superfluid density

ρs\displaystyle\rho_{s} =e2TNωn,kTr[G(k,iωn)vk,xG(k,iωn)vk,x+G(k,iωn)(xvk)τ3]\displaystyle=e^{2}\frac{T}{N}\sum_{\omega_{n},k}\text{Tr}\left[G(k,i\omega_{n})v_{k,x}G(k,i\omega_{n})v_{k,x}+G(k,i\omega_{n})(\nabla_{x}v_{k})\tau_{3}\right]{}
=4e2TNωn,kvk,x2Δk2[(|ωn|+Γk)2+εk2+Δk2]2\displaystyle=4e^{2}\frac{T}{N}\sum_{\omega_{n},k}\frac{v_{k,x}^{2}\Delta_{k}^{2}}{[(|\omega_{n}|+\Gamma_{k})^{2}+\varepsilon_{k}^{2}+\Delta_{k}^{2}]^{2}} (22)

where we keep only the xxxx-component. Similar to previous sections, the kk-summation is replaced by integrals over ε\varepsilon and θ\theta under wide band approximation with circular Fermi surface, giving rise to

ρs\displaystyle\rho_{s} =4e2T𝒩vF2ωndθ2πdεϕθ2Δθ2[(|ωn|+Γθ)2+ε2+Δθ2]2\displaystyle=4e^{2}T\mathcal{N}v_{F}^{2}\sum_{\omega_{n}}\int\frac{\mathrm{d}\theta}{2\pi}\int\mathrm{d}\varepsilon\frac{\phi_{\theta}^{2}\Delta_{\theta}^{2}}{[(|\omega_{n}|+\Gamma_{\theta})^{2}+\varepsilon^{2}+\Delta_{\theta}^{2}]^{2}}{}
=e2T𝒩vF2ωndθϕθ2Δθ2[(|ωn|+Γθ)2+Δθ2]3/2\displaystyle=e^{2}T\mathcal{N}v_{F}^{2}\sum_{\omega_{n}}\int\mathrm{d}\theta\frac{\phi_{\theta}^{2}\Delta_{\theta}^{2}}{[(|\omega_{n}|+\Gamma_{\theta})^{2}+\Delta_{\theta}^{2}]^{3/2}} (23)

which is Eq. 6 in the main text.

II.4 Optical conductivity

The optical conductivity is contributed only by the paramagnetic current as shown in Fig. 7.

{feynman}\vertexvkσ0v_{k}\sigma_{0}vkσ0v_{k}\sigma_{0}\vertex\diagramiνni\nu_{n}kiωn+iνnk\,i\omega_{n}+i\nu_{n}kiωnk\,i\omega_{n}iνni\nu_{n}
Figure 7: Feynman diagrams for the optical conductivity.

The current-current correlation function K(iνn)K(i\nu_{n}) is given by

K(iνn)\displaystyle K(i\nu_{n}) =e2TNωn,kTr[G(k,iωn)vk,xG(k,iωn+iνn)vkx]\displaystyle=-e^{2}\frac{T}{N}\sum_{\omega_{n},k}\text{Tr}\left[G(k,i\omega_{n})v_{k,x}G(k,i\omega_{n}+i\nu_{n})v_{k_{x}}\right]{}
=e2TNωn,kvk,x2dωdωTr[ImG(k,ω)ImG(k,ω)](iωnω)(iωn+iνnω)\displaystyle=-e^{2}\frac{T}{N}\sum_{\omega_{n},k}v_{k,x}^{2}\int\mathrm{d}\omega\int\mathrm{d}\omega^{\prime}\frac{\text{Tr}\left[\text{Im}G(k,\omega)\text{Im}G(k,\omega^{\prime})\right]}{(i\omega_{n}-\omega)(i\omega_{n}+i\nu_{n}-\omega^{\prime})} (24)

where the spectral representation has been used. Now the summation of iωni\omega_{n} can be completed and gives

K(iνn)=e2π2Nkvk,x2dωdωf(ω)f(ω)iνn+ωωTr[ImG(k,ω)ImG(k,ω)]\displaystyle K(i\nu_{n})=-\frac{e^{2}}{\pi^{2}N}\sum_{k}v_{k,x}^{2}\int\mathrm{d}\omega\int\mathrm{d}\omega^{\prime}\frac{f(\omega)-f(\omega^{\prime})}{i\nu_{n}+\omega-\omega^{\prime}}\text{Tr}\left[\text{Im}G(k,\omega)\text{Im}G(k,\omega^{\prime})\right] (25)

After analytic continuation iνnν+i0+i\nu_{n}\to\nu+i0^{+}, the optical conductivity σ1(ν)\sigma_{1}(\nu) can be obtained as

σ1(ν)\displaystyle\sigma_{1}(\nu) =Re[K(iνnν+i0+)iν]\displaystyle=\text{Re}\left[\frac{K(i\nu_{n}\to\nu+i0^{+})}{-i\nu}\right]{}
=e2π2Nkvk,x2dωf(ω)f(ω+ν)νTr[ImG(k,ω)ImG(k,ω+ν)]\displaystyle=\frac{e^{2}}{\pi^{2}N}\sum_{k}v_{k,x}^{2}\int\mathrm{d}\omega\frac{f(\omega)-f(\omega+\nu)}{\nu}\text{Tr}\left[\text{Im}G(k,\omega)\text{Im}G(k,\omega+\nu)\right] (26)

Finally, we replace the kk-summation by energy and angle integrals and arrive at the final expression

σ1(ν)=e2vF2𝒩2π3dεdωdθcos2θf(ω)f(ω+ν)νTr[ImG(ε,θ,ω)ImG(ε,θ,ω+ν)]\displaystyle\sigma_{1}(\nu)=\frac{e^{2}v_{F}^{2}\mathcal{N}}{2\pi^{3}}\int\mathrm{d}\varepsilon\int\mathrm{d}\omega\int\mathrm{d}\theta\cos^{2}\theta\frac{f(\omega)-f(\omega+\nu)}{\nu}\text{Tr}\left[\text{Im}G(\varepsilon,\theta,\omega)\text{Im}G(\varepsilon,\theta,\omega+\nu)\right] (27)

which is Eq. 7 in the main text.

Refer to caption
Figure 8: Fittings of the THz experimental data of σ1\sigma_{1} on four samples Mahmood et al. (2019) in the normal state (a) and superconducting state (b), respectively, with (solid lines) and without (dashed lines) the anisotropic scattering rate Γd\Gamma_{d}. In the superconducting state, we fix Δ0=2Tc\Delta_{0}=2T_{c} as a rough estimate of the pairing gap. For comparison, we also present the fitted results (dotted lines) with Δ0=0\Delta_{0}=0 as performed in Ref. Mahmood et al., 2019. In the inset of (a), we plot the best fitted parameters (Γs,Γd)(\Gamma_{s},\Gamma_{d}) versus TcT_{c}, respectively.

On the basis of Eq. 27, we fit the experimental data of σ1\sigma_{1} from Ref. Mahmood et al., 2019. We read the data points from ν=0.4\nu=0.4 to 1.61.6THz by every 0.10.1THz. The best fits are shown in Fig. 8(a) for normal state and in Fig. 8(b) for superconducting state, respectively, with (solid lines) and without (dashed or dotted lines) Γd\Gamma_{d}. For the superconducting state, we have fixed Δ0=2Tc\Delta_{0}=2T_{c} as a rough estimate of the pairing gap. Compared with pure Γs\Gamma_{s}-fitting, the results using both Gs and Gd agree better to most of the data. As anticipated, the pure-Γs\Gamma_{s} fitting underestimates both σ1(ν0)\sigma_{1}(\nu\to 0) and the spectral weight (enclosed area). We further show the best fitted parameters (Γs,Γd)(\Gamma_{s},\Gamma_{d}) obtained from the normal state data in the inset of Fig. 8(a), which shows clearly that Γd\Gamma_{d} becomes stronger upon overdoping (reducing TcT_{c}) while Γs\Gamma_{s} barely changes, supporting our main point that the apical oxygen vacancies become more and more important in overdoped LSCO upon doping.

III Further discussions

III.1 First principle calculations

In order to determine the position of the oxygen vacancies theoretically, we performed first principle calculations on La2-xSrxCuO4-y within the local density approximation (LDA), which has been found to work very well in the overdoped cuprates Kramer et al. (2019). By constructing two kinds of super unit cells, 2×2×12\times 2\times 1 and 2×2×22\times 2\times 2, with one oxygen vacancy, and optimizing the positions of the doped Sr atoms, we obtain the total energy difference between two kinds of oxygen vacancies ΔE=EapicalEplanar\Delta E=E_{\rm apical}-E_{\rm planar} as shown in Fig. 9(a). Without Sr doping, the planar oxygen vacancy EplanarE_{\rm planar} is lower than the apical one EapicalE_{\rm apical}, but upon Sr doping, their difference reduces and finally reverses at high doping levels. Such a tendency becomes stronger when we enlarge the super unit cell from 2×2×12\times 2\times 1 to 2×2×22\times 2\times 2, indicating the oxygen vacancies in overdoped LSCO are very likely to be apical rather than planar ones, in agreement with the Raman experiment Kim et al. (2017). This behavior can be roughly understood from the density of states (DOS), as shown in Fig. 9(b). As can be seen, the planner oxygen vacancy reduces the DOS near the Fermi level at x=0x=0 and thus saves more energy. Instead, the apical vacancy enhances the DOS near the Fermi level. Upon hole doping caused by Sr, the DOS of the apical case is reduced and thus finally becomes preferred than the planar case.

Refer to caption
Refer to caption
Figure 9: LDA results of apical and planar oxygen vacancies. (a) The energy difference ΔE=EapicalEplanar\Delta E=E_{\rm apical}-E_{\rm planar} is plotted versus the Sr concentration (2x) with two kinds of super unit cells. (b) The total DOS near the Fermi level at x=0x=0 for apical and planar oxygen vacancies, respectively.

III.2 Planar oxygen vacancies

From both the experiment and our LDA calculations, we have found the oxygen vacancies are dominated by apical ones in overdoped LSCO. However, at small doping, the planar ones are more stable. Therefore, we may also ask what happens for the planar oxygen vacancies?

At first, one planar oxygen vacancy breaks the Zhang-Rice singlets on its two neighboring sites. To describe such an effect, we consider a model by adding two onsite potentials given by

Hx,yI=V(d0d0+dx,ydx,y)=VNkk[1+ei(kx,ykx,y)]dkdk\displaystyle H^{\rm I}_{x,y}=V(d_{0}^{\dagger}d_{0}+d_{x,y}^{\dagger}d_{x,y})=\frac{V}{N}\sum_{kk^{\prime}}\left[1+\mathrm{e}^{i(k_{x,y}^{\prime}-k_{x,y})}\right]d_{k}^{\dagger}d_{k^{\prime}} (28)

where the subscript x,yx,y labels the position of the vacancy is on the xx-bond or yy-bond, hence, d0d_{0}^{\dagger}, dxd_{x}^{\dagger} and dyd_{y}^{\dagger} generate electrons on (0,0)(0,0), (x^,0)(\hat{x},0) and (y^,0)(\hat{y},0), respectively. We call this scattering process as type-I. Its amplitude |Vkk|2|V_{kk^{\prime}}|^{2} is given by

|VkkI|2=4V2cos2(kx,ykx,y2)\displaystyle|V^{\rm I}_{kk^{\prime}}|^{2}=4V^{2}\cos^{2}\left(\frac{k_{x,y}^{\prime}-k_{x,y}}{2}\right) (29)

Such a disorder scattering is difficult to treat exactly as it depends on the transferred momentum (kx,ykx,y)(k_{x,y}-k_{x,y}^{\prime}) (but not kk and kk^{\prime} independently like the apical case) and breaks the rotational symmetry. Nevertheless, let us qualitatively consider its role on the d-wave superconducting properties. If we keep only the dominant scattering processes, they occur at kxkxk_{x}\approx k_{x}^{\prime} (for xx-bond vacancies)and kykyk_{y}\approx k_{y}^{\prime} (for yy-bond vacancies). These scatterings preserve the pairing sign and are expected to be not pair-breaking, in some sense like the point disorder in s-wave superconductors. Therefore, the standard nonmagnetic disorder effects on the conventional dirty s-wave superconductors are expected: such disorders should have little effect on TcT_{c} but reduce the superfluid density. Furthermore, these scatterings do not vanish for nodal quasiparticles, hence, the scattering rate Γ\Gamma is finite in this region (like Γs\Gamma_{s}) which cannot explain the linear TT-dependence of the superfluid density.

On the other hand, one planar oxygen vacancy can also change the hopping on this bond (called type-II process), which can be described by the following Hamiltonian

Hx,yII=V(d0dx,y+dx,yd0)=VNkk(eikx,y+eikx,y)dkdk\displaystyle H^{\rm II}_{x,y}=V^{\prime}(d_{0}^{\dagger}d_{x,y}+d_{x,y}^{\dagger}d_{0})=\frac{V^{\prime}}{N}\sum_{kk^{\prime}}\left(\mathrm{e}^{ik^{\prime}_{x,y}}+\mathrm{e}^{-ik_{x,y}}\right)d_{k}^{\dagger}d_{k^{\prime}} (30)

The symbols are similar to the above. The amplitude of this scattering potential is

|VkkII|2=4V2cos2(kx,y+kx,y2)\displaystyle|V^{\rm II}_{kk^{\prime}}|^{2}=4V^{\prime 2}\cos^{2}\left(\frac{k_{x,y}^{\prime}+k_{x,y}}{2}\right) (31)

which is largest at kxkxk_{x}\approx-k_{x}^{\prime} (for xx-bond vacancies) and kykyk_{y}\approx-k_{y}^{\prime} (for yy-bond vacancies). These scatterings are also not pair-breaking and thus are expected to be similar to the above type-I scatterings.

Put these two processes together, we conclude that the planar oxygen vacancies behave more like pair-preserving disorders in d-wave superconductors and cannot explain the superfluid density experiment.

III.3 Current vertex correction

Now let us examine the effect of the current vertex correction caused by the anisotropic disorder scattering. We label the bare current vertex by vkv_{k} and the dressed vertex by Λk\Lambda_{k}. They are connected in a self-consistent way as shown in Fig. 10, where we have neglected all crossing diagrams which are at least of order nimp2κ4n_{\rm imp}^{2}\kappa^{4} and are expected to be more relevant to localization Lee and Ramakrishnan (1985). In fact, the resistivity in the overdoped LSCO becomes smaller and smaller upon doping, indicating the system are more and more itinerant than localization Božović et al. (2016).

{feynman}\vertex\vertex\vertexΛk\vertex\diagramkk={feynman}\vertex\vertex\vertexvk\vertex\diagramkk+{feynman}\vertex\vertexnimpΛk\vertex\vertex\vertex\vertex\vertex\diagramkkkkVkkVkk\displaystyle\vbox{\hbox{ \leavevmode\hbox to20.86pt{\vbox to22.51pt{\pgfpicture\makeatletter\hbox{\hskip-42.87912pt\lower-3.33301pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\feynman \vertex(i); \vertex[below=3cm of i] (j); \vertex[below=1.5cm of i] (k); {{}}{{}} {{{}{{}{}}{}}}{{ {}{}{}}}{{{{}}{{}}}}{{}}{{{ }}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{46.21213pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{}}{} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{50.3681pt}{9.01044pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$\Lambda_{k}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \vertex[right=1.5cm of v] (o); \diagram*{ (i) --[fermion, edge label=$k$] (v) -- [fermion, edge label=$k$] (j), (v) --[photon] (o), }; \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}}~{}=~{}\vbox{\hbox{ \leavevmode\hbox to19.12pt{\vbox to19.98pt{\pgfpicture\makeatletter\hbox{\hskip-42.87912pt\lower-3.33301pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\feynman \vertex(i); \vertex[below=3cm of i] (j); \vertex[below=1.5cm of i] (k); {{}}{{}} {{{}{{}{}}{}}}{{ {}{}{}}}{{{{}}{{}}}}{{}}{{{ }}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{46.21213pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{}}{} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{50.3681pt}{9.01044pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$v_{k}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \vertex[right=1.5cm of v] (o); \diagram*{ (i) --[fermion,edge label=$k$] (v) -- [fermion,edge label=$k$] (j), (v) --[photon] (o), }; \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}}~{}+~{}\vbox{\hbox{ \leavevmode\hbox to90.35pt{\vbox to68.53pt{\pgfpicture\makeatletter\hbox{\hskip 25.73471pt\lower-48.90944pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\feynman \vertex(i); \vertex[below=3cm of i] (j); {{}}{{}} {{{}{{}{}}{}}}{{ {}{}{}}}{{{{}}{{}}}}{{}}{{{ }}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{-41.94423pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{{}{}}}{{}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-22.4017pt}{-42.61757pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$n_{\rm imp}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}{{}} {{{}{{}{}}{}}}{{ {}{}{}}}{{{{}}{{}}}}{{}}{{{ }}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{46.21213pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{}}{} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{50.3681pt}{9.4549pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$\Lambda_{k^{\prime}}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \vertex[right=1.5cm of v] (o); \vertex[below=0.5cm of i] (x); \vertex[right=0.5cm of x] (i2); \vertex[above=0.5cm of j] (y); \vertex[right=0.5cm of y] (j2); \diagram*{ (i) --[fermion, edge label=$k$] (i2) -- [fermion, edge label=$k^{\prime}$] (v) -- [fermion, edge label=$k^{\prime}$] (j2) --[fermion, edge label=$k$] (j), (v) --[photon] (o), (j2) -- [scalar,edge label=$V_{kk^{\prime}}$] (k) -- [scalar, edge label=$V_{k^{\prime}k}$] (i2), }; \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}}{}
Figure 10: Current vertex correction caused by the disorder scattering.

From Fig. 10, we can write down the self-consistent equation

Λk=vk+nimpNk|Vkk|2G2(iωn,k)Λk\displaystyle\Lambda_{k}=v_{k}+\frac{n_{\rm imp}}{N}\sum_{k^{\prime}}|V_{kk^{\prime}}|^{2}G^{2}(i\omega_{n},k^{\prime})\Lambda_{k^{\prime}} (32)

For a general scattering VkkV_{kk^{\prime}}, the above equation can be solved to give rise to an additional factor (1cosθkk)(1-\cos\theta_{kk^{\prime}}) in the so-called transport scattering rate τtr1\tau_{\rm tr}^{-1} which enters σ1(ν)\sigma_{1}(\nu) Abrikosov et al. (1963). But here, in our model, it becomes quite simple because the disorder scattering potential VkkV_{kk^{\prime}} is already factorized Vkk=16κcos2θkcos2θkV_{kk^{\prime}}=16\kappa\cos 2\theta_{k}\cos 2\theta_{k^{\prime}} such that the summation over kk^{\prime} can be completed exactly. Moreover, notice that Λk\Lambda_{k^{\prime}} is time reversal (kkk^{\prime}\to-k^{\prime}) odd, while |Vkk|2G2(iωn,k)|V_{kk^{\prime}}|^{2}G^{2}(i\omega_{n},k^{\prime}) is time reversal even, the integral over kk^{\prime} must be zero, which means Λk=vk\Lambda_{k}=v_{k}. Therefore, in our disorder model, the current vertex correction totally vanishes, if the crossing diagrams can be safely neglected.