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

Scattering of electron from a disk in 2D electron gas: full cross section, transport cross section, and the interaction correction

Nathan L. Foulk    M. E. Raikh Department of Physics and Astronomy, University of Utah, Salt Lake City, UT 84112
Abstract

It is known that the presence of the Fermi sea modifies the scattering of an electron from a point-like impurity. This is due to the Friedel oscillations of the electron density around the impurity. These oscillations create an additional scattering potential for incident electrons. The closer the energy of the incident electron to the Fermi level, the stronger the additional scattering. We study this effect for the case when the impurity is not point-like but rather a hard disk, with a radius much bigger than the de Broglie wavelength. We start with a careful examination of the full and transport cross sections from an extended target. Both cross sections approach their limiting values upon increasing the wave vector of the incident electron. We establish that the transport cross section saturates much faster than the full cross section. With regard to the interaction correction, we establish that it vanishes for the full cross section, while for the transport cross section, it is enhanced compared to the case of a point-like scatterer.

I Introduction

The static polarization operator, Π(𝐪)\Pi({\bf q}), of the 2D electron gas contains a singular correction,stern1967 Mπ2(q2kF)1/2\frac{M}{\pi\hbar^{2}}\left(q-2k_{\scriptscriptstyle F}\right)^{1/2}, near q=2kFq=2k_{\scriptscriptstyle F}. Here MM is the electron mass and kFk_{\scriptscriptstyle F} is the Fermi momentum. This singular behavior (Kohn anomaly) translates into the interaction corrections to the thermodynamic characteristics of the 2D gas,Chubukov such as effective mass. These corrections exhibit singular behavior as a function of temperature, TT. Transport characteristics of the 2D gas, such as conductivity, also acquire singular interaction corrections in the ballistic regime, Tτ1T\tau\gg 1, where τ\tau is the scattering time. In this regime, multiple scattering of electrons by the impurities can be neglected, while modification of the potential of individual impurities due to the Kohn anomalyGold1986 yields a correction to the scattering cross section proportional to TT. This mechanism of the anomalous temperature dependence was pointed out in Ref. Gold1986, . It was subsequently elaborated upon in Refs. Rudin1997, ; Zala, ; Gornyi2006, . Consideration of Refs.  Rudin1997, ; Zala, ; Gornyi2006, led to the following lucid prescription for incorporating electron-electron interactions into the calculation of transport.

An impurity potential, Uimp(𝐫)U_{\text{imp}}({\bf r}), causes a perturbation of the electron density

n(𝐫)n0=ν0sin(2kFr)2πr2𝑑𝐫Uimp(𝐫),n({\bf r})-n_{0}=-\nu_{0}\frac{\sin(2k_{\scriptscriptstyle F}r)}{2\pi r^{2}}\int d{\bf r}U_{\text{imp}}({\bf r}), (1)

around it. Here ν0=Mπ2\nu_{0}=\frac{M}{\pi\hbar^{2}} is the density of states. The Friedel oscillations Eq. (1) translate into an additional scattering potential for incident electrons. The correction to the scattering amplitude due to this potential depends dramatically on the energy, ε\varepsilon, of the incident electron measured from the Fermi level, EFE_{\scriptscriptstyle F}. Perturbative calculation of this correctionZala indicates that it is maximal within the angular interval |εEF1|1/2\sim\Big{|}\frac{\varepsilon}{E_{\scriptscriptstyle F}}-1\Big{|}^{1/2} near the backscattering condition. The relative magnitude of this correction is also |εEF1|1/2\sim\Big{|}\frac{\varepsilon}{E_{\scriptscriptstyle F}}-1\Big{|}^{1/2}. Thus, for a typical value |εEF|T|\varepsilon-E_{\scriptscriptstyle F}|\sim T, the relative interaction correction to the net scattering cross section can be estimated as TEF\frac{T}{E_{\scriptscriptstyle F}}.

The calculations in Refs. Rudin1997, ; Zala, ; Gornyi2006, were carried out for point-like scatterers. Namely, it was assumed that their size is much smaller than the de Broglie wave length, 2πkF\frac{2\pi}{k_{\scriptscriptstyle F}}. In the present paper, we extend the theoryRudin1997 ; Zala ; Gornyi2006 to the scatterers of the arbitrary size, aa. For this purpose, we first analyze the Friedel oscillations from the disk and also study the behavior of the full and transport cross sections from the disk in the absence of interactions. Then we incorporate interactions and study corresponding corrections to these cross sections.

II Friedel oscillations from the wall

In terms of formation of the Friedel oscillations, at small distances, (ra)a(r-a)\ll a, the scatterer can be viewed as a hard wall. Then the electron density does not depend on aa. To calculate this density it is sufficient to substitute into the definition

n(𝐫)=𝐤Θ(EF2𝐤22M)|Ψ𝐤(𝐫)|2,n({\bf r})=\sum_{\bf k}\Theta\Bigl{(}E_{\scriptscriptstyle F}-\frac{\hbar^{2}{\bf k}^{2}}{2M}\Bigr{)}|\Psi_{\bf k}({\bf r})|^{2}, (2)

the wave functions, Ψ𝐤=eikyysinkxx\Psi_{\bf k}=e^{ik_{y}y}\sin k_{x}x, which turn to zero at the wall. Here Θ(z)\Theta(z) is a step-function. Performing the integration over the components of the wave vector, we obtain

n(x)n0=n0J1(2kFx)kFx,n(x)-n_{0}=-n_{0}\frac{J_{1}(2k_{\scriptscriptstyle F}x)}{k_{\scriptscriptstyle F}x}, (3)

where J1(z)J_{1}(z) is the first-order Bessel function. These oscillations are shown in Fig. 1. The concentration is zero at x=0x=0. At large distances, kFx1k_{\scriptscriptstyle F}x\gg 1, the relative correction to the density is small and falls off as (kFx)3/2\left(k_{\scriptscriptstyle F}x\right)^{-3/2}, which is intermediate between (kFx)2\left(k_{\scriptscriptstyle F}x\right)^{-2} in 2D and (kFx)1\left(k_{\scriptscriptstyle F}x\right)^{-1} in 1D.

Refer to caption
Figure 1: (Color online) (a) blue curve: Friedel oscillations from the wall are plotted from Eq. (3); yellow curve: Friedel oscillations from the disk are plotted from Eq. (9) for kFa=3k_{F}a=3. (b) same dependencies as in (a) are plotted for large values of 2kFx2k_{\scriptscriptstyle F}x. Oscillations from the disk fall of faster than the oscillations from the wall.

II.1 Friedel oscillations from a hard disk

To calculate the radial dependence of the electron density,

n(r)k,mΘ(EF2k22M)[Rm,k(0)(r)]2,n(r)\propto\sum_{k,m}\Theta\left(E_{\scriptscriptstyle F}-\frac{\hbar^{2}k^{2}}{2M}\right)\left[R_{m,k}^{(0)}(r)\right]^{2}, (4)

we use the semiclassical form of the eigenfunctions

Rm,k(0)(r)=sin[ar𝑑r(k2m21/4r2)1/2](k2r2m2+14)1/4.R_{m,k}^{(0)}(r)=\frac{\sin\Bigl{[}\int\limits_{a}^{r}dr^{\prime}\left(k^{2}-\frac{m^{2}-1/4}{r^{\prime 2}}\right)^{1/2}\Bigr{]}}{\left(k^{2}r^{2}-m^{2}+\frac{1}{4}\right)^{1/4}}. (5)

The oscillating part of n(r)n(r) is determined by the states with energies close to the Fermi level, i.e. with kk close to kFk_{\scriptscriptstyle F}. This allows to expand the argument of sine in Rm,k(0)R_{m,k}^{(0)} with respect to δk=kkF\delta k=k-k_{\scriptscriptstyle F} as follows

(k2m214r2)1/2=(kF2m214r2)1/2kFδk(kF2m214r2)1/2.\left(k^{2}-\frac{m^{2}-\frac{1}{4}}{r^{\prime 2}}\right)^{1/2}\!\!\\ =\left(k_{\scriptscriptstyle F}^{2}-\frac{m^{2}-\frac{1}{4}}{r^{\prime 2}}\right)^{1/2}\!\!-\frac{k_{\scriptscriptstyle F}\delta k}{\left(k_{\scriptscriptstyle F}^{2}-\frac{m^{2}-\frac{1}{4}}{r^{\prime 2}}\right)^{1/2}}. (6)

It is also sufficient to set k=kFk=k_{\scriptscriptstyle F} in the prefactor of Rm,k(0)R_{m,k}^{(0)}. Then the integration over δk\delta k yields

m=\displaystyle\sum_{m=-\infty}^{\infty} {1+(kF2a2m2kF2r2m2)1/2}\displaystyle\Biggl{\{}1+\Biggl{(}\frac{k_{\scriptscriptstyle F}^{2}a^{2}-m^{2}}{k_{\scriptscriptstyle F}^{2}r^{2}-m^{2}}\Biggr{)}^{1/2}\Biggr{\}}
×sin[2ar𝑑r(kF2m2r2)1/2]kF2(r2a2).\displaystyle\times\frac{\sin\Bigl{[}2\int\limits_{a}^{r}dr^{\prime}\left(k_{\scriptscriptstyle F}^{2}-\frac{m^{2}}{r^{\prime 2}}\right)^{1/2}\Bigr{]}}{k_{\scriptscriptstyle F}^{2}\left(r^{2}-a^{2}\right)}. (7)

The expansion Eq. (6) is justified for δkkF\delta k\ll k_{\scriptscriptstyle F}. On the other hand, characteristic δk\delta k in the integration is δk(ra)1\delta k\sim\left(r-a\right)^{-1}. Thus, the result Eq. (7) applies for distances (ra)kF1(r-a)\gg k_{\scriptscriptstyle F}^{-1}. This result describes the oscillating part of n(r)n(r) within a numerical factor.

As we will see later, the main contribution to the sum Eq. (7) comes from large momenta, m1m\gg 1, but still with mkam\ll ka. In this domain, one can expand the integrand in Eq. (7) and perform the integration over drdr^{\prime}, which yields

ar𝑑r(kF2m2r2)1/2kF(ra)m22kF(1a1r).\int\limits_{a}^{r}dr^{\prime}\left(k_{\scriptscriptstyle F}^{2}-\frac{m^{2}}{r^{\prime 2}}\right)^{1/2}\approx k_{\scriptscriptstyle F}(r-a)-\frac{m^{2}}{2k_{\scriptscriptstyle F}}\left(\frac{1}{a}-\frac{1}{r}\right). (8)

We see that the condition mkFam\ll k_{\scriptscriptstyle F}a ensures that the second term in Eq. (8) is much smaller than the first term. On the other hand, the second term in Eq. (8) defines characteristic m(kFarra)1/2m\sim\left(\frac{k_{\scriptscriptstyle F}ar}{r-a}\right)^{1/2}. This value is smaller than kFak_{\scriptscriptstyle F}a under the condition (ra)kF1\left(r-a\right)\gg k_{\scriptscriptstyle F}^{-1}, which we have already assumed to be met. Still, this value is much bigger than one, which allows us to replace the summation over mm by the integration. The final result for the oscillating part of n(r)n(r) reads

δn(r)n0=(32arπ)1/2(1r+a)sin[2kF(ra)π4][2kF(ra)]3/2.\frac{\delta n(r)}{n_{0}}=-\left(\frac{32ar}{\pi}\right)^{1/2}\!\!\!\left(\frac{1}{r+a}\right)\frac{\sin\Bigl{[}2k_{\scriptscriptstyle F}(r-a)-\frac{\pi}{4}\Bigr{]}}{\Bigl{[}2k_{\scriptscriptstyle F}(r-a)\Bigr{]}^{3/2}}. (9)

In Eq. (9) we have restored the numerical factor. In the domain kF1(ra)ak_{\scriptscriptstyle F}^{-1}\ll(r-a)\ll a, Eq. (9) reproduces the Friedel oscillations Eq. (3) from a hard wall, while for (ra)a(r-a)\gg a the oscillations fall off as 1/r21/r^{2} like for a point-like impurity.

III Scattering in the absence of interactions

III.1 Basic relations

The general expression for the scattering cross section reads

σ=02π𝑑φ|f(φ)|2,\sigma=\int\limits_{0}^{2\pi}d\varphi|f(\varphi)|^{2}, (10)

where f(φ)f(\varphi) is the scattering amplitude. It is related to the scattering phases, δm(0)\delta_{m}^{(0)}, as followsLandau

f(φ)=1(2πik)1/2m=(e2iδm(0)1)eimφ.f(\varphi)=\frac{1}{\left(2\pi ik\right)^{1/2}}\sum_{m=-\infty}^{\infty}\left(e^{2i\delta_{m}^{(0)}}-1\right)e^{im\varphi}. (11)

Scattering phases are determined by the long-distance behavior of the radial wave functions

Rm,k(r)|r1r1/2cos(krmπ2π4+δm(0)),R_{m,k}(r)\Big{|}_{r\rightarrow\infty}\propto\frac{1}{r^{1/2}}\cos\left(kr-\frac{m\pi}{2}-\frac{\pi}{4}+\delta_{m}^{(0)}\right), (12)

where k=1(2mE)1/2k=\frac{1}{\hbar}\left(2mE\right)^{1/2} is the wave vector. Performing the angular integration, one recovers the textbook resultLandau

σ=4km=sin2δm(0).\sigma=\frac{4}{k}\sum_{m=-\infty}^{\infty}\sin^{2}\delta_{m}^{(0)}. (13)

The quantity that enters the conductivity is not σ\sigma but the transport cross section defined as

σtr=02π𝑑φ(1cosφ)|f(φ)|2.\sigma_{\text{tr}}=\int\limits_{0}^{2\pi}d\varphi(1-\cos\varphi)|f(\varphi)|^{2}. (14)

Substituting Eq. (11) into Eq. (14) and integrating over φ\varphi, one obtains

σtr=2km=sin2(δm(0)δm+1(0)).\sigma_{\text{tr}}=\frac{2}{k}\sum_{m=-\infty}^{\infty}\sin^{2}\left(\delta_{m}^{(0)}-\delta_{m+1}^{(0)}\right). (15)

IV Hard disk

IV.1 Classical calculation

Refer to caption
Figure 2: (Color online) A schematic of classical hard-disk scattering. A particle with impact parameter bb incident on a target with radius aa. The scattering angle θ\theta is related to the impact parameter bb through Eq. (16).

Within a classical picture, a particle incident on the disk with impact parameter, b<ab<a, is reflected from the boundary, see Fig. 2. The scattering angle, θ\theta, is related to bb as

b(θ)=acos(θ2).b(\theta)=-a\cos\left(\frac{\theta}{2}\right). (16)

Substituting Eq. (16) into the definitions of the full and transport cross sections

σ=02π𝑑θ|bθ|,σtr=02π𝑑θ(1cosθ)|bθ|,\sigma=\int\limits_{0}^{2\pi}d\theta\Big{|}\frac{\partial b}{\partial\theta}\Big{|},~{}~{}~{}~{}\sigma_{\text{tr}}=\int\limits_{0}^{2\pi}d\theta\left(1-\cos\theta\right)\Big{|}\frac{\partial b}{\partial\theta}\Big{|}, (17)

we get

σ=a202π𝑑θsin(θ2)=2a,\displaystyle\sigma=\frac{a}{2}\int\limits_{0}^{2\pi}d\theta\sin\left(\frac{\theta}{2}\right)=2a,
σtr=a202π𝑑θ(1cosθ)sin(θ2)=83a.\displaystyle\sigma_{\text{tr}}=\frac{a}{2}\int\limits_{0}^{2\pi}d\theta\left(1-\cos\theta\right)\sin\left(\frac{\theta}{2}\right)=\frac{8}{3}a. (18)

The above calculation suggests that the classical cross section is equal to the diameter of the disk. Although this diameter is much bigger than the de Broglie wave length, the result obtained neglecting the diffraction effects is not supported by the quantum calculation.

IV.2 Quantum calculation

For a hard disk of a radius, aa, the form of the radial wave function at r>ar>a is the linear combination

Rm,k(r)=cosδmJm(kr)+sinδmNm(kr),R_{m,k}(r)=\cos\delta_{m}J_{m}(kr)+\sin\delta_{m}N_{m}(kr), (19)

of the Bessel and the Neumann functions, Jm(z)J_{m}(z) and Nm(z)N_{m}(z), which are the free solutions of the Schrödinger equation. Then the exact expression for the phases, δm\delta_{m}, which follows from the condition Rm,k(a)=0R_{m,k}(a)=0, reads

sin2δm(0)=Jm2(ka)Jm2(ka)+Nm2(ka).\sin^{2}\delta_{m}^{(0)}=\frac{J_{m}^{2}(ka)}{J_{m}^{2}(ka)+N_{m}^{2}(ka)}. (20)

At small ka1ka\ll 1 the sum Eq. (13) is dominated by the first term for which J0(ka)1J_{0}(ka)\approx 1 and N0(ka)2πln(ka)N_{0}(ka)\approx\frac{2}{\pi}\ln(ka). Then Eq. (13) takes the form

σ4a|ka1π24kaln2(ka).\frac{\sigma}{4a}\Big{|}_{ka\ll 1}\approx\frac{\pi^{2}}{4ka\ln^{2}(ka)}. (21)

The right-hand side has a minimum at ka0.25ka\approx 0.25. At this kaka the terms with higher mm in Eq. (13) are important, leading to the decay of σ\sigma with energy. Still, the minimum in m=0m=0 term manifests itself as a shallow minimum in the derivative dσd(ka)\frac{d\sigma}{d(ka)}, as illustrated by the numerical calculation, see Fig. 3.

Upon increasing energy the fall off of the cross section saturates at σ4a1\frac{\sigma}{4a}\approx 1, which corresponds to replacement of sin2δm\sin^{2}\delta_{m} by 1/21/2 for m<kam<ka and by zero for m>kam>ka. This value exceeds twice the classical result Eq. (IV.1), which is a well-known effect in 3D, see e.g. Ref. Landau, .

Next we will study the quantum correction to the scattering cross section, which determines the law of approach of σ\sigma to the saturation value in the limit ka1ka\gg 1. The main point is that this approach is dominated by a narrow domain of momenta (kam)ka\left(ka-m\right)\ll ka. To capture the quantum correction analytically, we infer the expression for the phases by comparing the semiclassical form Eq. (5) of the radial wave functions to the asymptote Eq. (12). The integral in the argument of sine can be evaluated analytically

Im\displaystyle I_{m}\! =ar(k2m2r2)1/2𝑑r\displaystyle=\int\limits_{a}^{r}\left(k^{2}-\frac{m^{2}}{{r^{\prime}}^{2}}\right)^{1/2}\!dr^{\prime}
=(k2r2m2)1/2(k2a2m2)1/2\displaystyle=\left(k^{2}r^{2}-m^{2}\right)^{1/2}-\left(k^{2}a^{2}-m^{2}\right)^{1/2}
marctan[(krm)21]1/2+marctan[(kam)21]1/2.\displaystyle-m\arctan\Bigl{[}\left(\frac{kr}{m}\right)^{2}-1\Bigr{]}^{1/2}\!\!\!+m\arctan\Bigl{[}\left(\frac{ka}{m}\right)^{2}-1\Bigr{]}^{1/2}\!\!\!. (22)

Taking the limit rr\rightarrow\infty we obtain

Im\displaystyle I_{m} krmπ2[k2a2m2]1/2\displaystyle\approx kr-\frac{m\pi}{2}-\left[k^{2}a^{2}-m^{2}\right]^{1/2}
+marctan[(kam)21]1/2.\displaystyle+m\arctan\left[\left(\frac{ka}{m}\right)^{2}-1\right]^{1/2}. (23)

With this ImI_{m}, the semiclassical expression Eq. (5) matches the asymptote Eq. (12) for the following choice of the scattering phases

3π4δm(0)\displaystyle\frac{3\pi}{4}-\delta_{m}^{(0)}
=m{[(kam)21]1/2arctan[(kam)21]1/2}.\displaystyle=m\Biggr{\{}\left[\left(\frac{ka}{m}\right)^{2}-1\right]^{1/2}\mkern-18.0mu-\arctan\left[\left(\frac{ka}{m}\right)^{2}-1\right]^{1/2}\Biggr{\}}. (24)

In the domain (kam)ka\left(ka-m\right)\ll ka the argument of arctangent is small, which allows to use the expansion zarctanz=13z3z-\arctan z=\frac{1}{3}z^{3}, after which Eq. (IV.2) simplifies to

3π4δm(0)(2m1)3/23(ka)1/2,\frac{3\pi}{4}-\delta_{m}^{(0)}\approx\frac{(2m_{1})^{3/2}}{3(ka)^{1/2}}, (25)

where m1=kamm_{1}=ka-m characterizes the proximity of angular momentum to kaka. From here we get

sin2δm(0)=12+12sin[2(2m1)3/23(ka)1/2].\sin^{2}\delta_{m}^{(0)}=\frac{1}{2}+\frac{1}{2}\sin\Biggl{[}\frac{2(2m_{1})^{3/2}}{3(ka)^{1/2}}\Biggr{]}. (26)

The argument of sine defines the characteristic m1(ka)1/3m_{1}\sim(ka)^{1/3}, which is much smaller than kaka, as was assumed above. On the other hand, in the domain ka1ka\gg 1, this characteristic value is much bigger than 11, which allows one to replace the summation over m1m_{1} by integration. This yields

σ4a1\displaystyle\frac{\sigma}{4a}-1 =0dm1kasin[2(2m1)3/23(ka)1/2]\displaystyle=\int\limits_{0}^{\infty}\frac{dm_{1}}{ka}\sin\Biggl{[}\frac{2(2m_{1})^{3/2}}{3(ka)^{1/2}}\Biggr{]}
=12(23k2a2)1/30dzsinzz1/3\displaystyle=\frac{1}{2}\left(\frac{2}{3k^{2}a^{2}}\right)^{1/3}\!\!\int\limits_{0}^{\infty}\frac{dz\sin z}{z^{1/3}}
=πΓ(13)1121/31(ka)2/3=α(ka)2/3,\displaystyle=\frac{\pi}{\Gamma(\frac{1}{3})}\frac{1}{12^{1/3}}\frac{1}{(ka)^{2/3}}=\frac{\alpha}{(ka)^{2/3}}, (27)

where α=πΓ(13)1121/30.51\alpha=\frac{\pi}{\Gamma(\frac{1}{3})}\frac{1}{12^{1/3}}\approx 0.51.

The right-hand side of Eq. (IV.2) is the amount by which the cross section at finite energy exceeds the limiting value, σ=4a\sigma=4a.

Refer to caption
Figure 3: (Color online) Full scattering cross section is calculated numerically from Eqs. (13) and (20). In (a), the approach to the asymptotic value σ=4a\sigma=4a is shown. In (b), the log-log plot of σ(ka)\sigma(ka) calculated numerically (blue curve) is compared to the theoretical prediction Eq. (IV.2) (yellow curve). In (c), the slope dσdlnka\frac{d\sigma}{d\ln ka} is plotted versus kaka. In agreement with theory, Eq. (IV.2) the slope approaches 2/32/3 at large kaka. The minimum at ka0.2ka\approx 0.2 originates from m=0m=0 term, Eq. (21).

IV.3 Transport cross section in 2D

The contribution to σ\sigma and to σtr\sigma_{\text{tr}} from big momenta, |m|>ka|m|>ka, is exponentially small. For |m|ka|m|\ll ka the asymptotic expression for the phases isLapidus

δm(0)=arctan(Jm(ka)Nm(ka))kaπ2(m+1)π4.\delta_{m}^{(0)}=-\arctan\Biggl{(}\frac{J_{m}(ka)}{N_{m}(ka)}\Biggr{)}\approx ka-\frac{\pi}{2}\left(m+1\right)-\frac{\pi}{4}. (28)

A more general expression for δm(0)\delta_{m}^{(0)} which is valid for all mm smaller than kaka can be inferred from the semiclassical expression Eq. (5) by calculating the integral in the argument of sine explicitly. This yields

δm(0)=[(ka)2m2]1/2marctan[(kam)21]1/2+3π4.\delta_{m}^{(0)}=\left[\left(ka\right)^{2}-m^{2}\right]^{1/2}-m\arctan\left[\left(\frac{ka}{m}\right)^{2}-1\right]^{1/2}+\frac{3\pi}{4}. (29)

In calculating the full cross section it is sufficient to replace sin2δm(0)\sin^{2}\delta_{m}^{(0)} by 1/21/2 in all 2ka+12ka+1 terms in Eq. (13) for which the phase is big. Upon doing so, the standard resultLapidus

σ4a,\sigma\approx 4a, (30)

is reproduced. As in the 3D case,Landau the result Eq. (30) does not contain the wavelength of the incident electron. Still, the full cross section exceeds twice the geometric cross section.

The calculation of the transport cross section is a more delicate task, since the difference δm(0)δm+1(0)\delta_{m}^{(0)}-\delta_{m+1}^{(0)} is a slow function of mm. Indeed, from Eq. (29) we get

δm(0)δm+1(0)δm(0)m=arctan(k2a2m21)1/2.\delta_{m}^{(0)}-\delta_{m+1}^{(0)}\approx-\frac{\partial\delta_{m}^{(0)}}{\partial m}=\arctan\left(\frac{k^{2}a^{2}}{m^{2}}-1\right)^{1/2}. (31)

From the above relation we find

sin2(δm(0)δm+1(0))1(mka)2.\sin^{2}\Bigl{(}\delta_{m}^{(0)}-\delta_{m+1}^{(0)}\Bigr{)}\approx 1-\Bigl{(}\frac{m}{ka}\Bigr{)}^{2}. (32)

Then the summation over mm leads to the following result for the transport scattering cross section

σtr83a.\sigma_{\text{tr}}\approx\frac{8}{3}a. (33)

Unlike the full cross section, this result coincides coincides with the transport cross section calculated classically. To study the quantum correction to σtr\sigma_{\text{tr}}, we express the difference δm(0)δm+1(0)\delta_{m}^{(0)}-\delta_{m+1}^{(0)} in terms of the Bessel functions. This yields

σtr4a\displaystyle\mkern-18.0mu\frac{\sigma_{\text{tr}}}{4a} =12kam=sin2(δm(0)δm+1(0))\displaystyle=\frac{1}{2ka}\sum_{m=-\infty}^{\infty}\sin^{2}\left(\delta_{m}^{(0)}-\delta_{m+1}^{(0)}\right)
=12kam=[Jm(ka)Nm+1(ka)Nm(ka)Jm+1(ka)]2[Jm2(ka)+Nm2(ka)][Jm+12(ka)+Nm+12(ka)].\displaystyle=\frac{1}{2ka}\!\!\!\sum_{m=-\infty}^{\infty}\!\!\frac{\left[J_{m}(ka)N_{m+1}(ka)-N_{m}(ka)J_{m+1}(ka)\right]^{2}}{\left[J^{2}_{m}(ka)+N^{2}_{m}(ka)\right]\left[J_{m+1}^{2}(ka)+N_{m+1}^{2}(ka)\right]}. (34)

The numerator of Eq. (34) can be greatly simplified upon using the relationBateman

Jm(ka)Nm+1(ka)Nm(ka)Jm+1(ka)=2πka.J_{m}(ka)N_{m+1}(ka)-N_{m}(ka)J_{m+1}(ka)=-\frac{2}{\pi ka}. (35)

With this simplification Eq. (34) assumes the form

σtr4a=4π2(ka)3×m=0(1[Jm2(ka)+Nm2(ka)][Jm+12(ka)+Nm+12(ka)]).\frac{\sigma_{\text{tr}}}{4a}=\frac{4}{\pi^{2}(ka)^{3}}\\ \times\sum_{m=0}^{\infty}\Biggr{(}\frac{1}{\left[J^{2}_{m}(ka)+N^{2}_{m}(ka)\right]\left[J_{m+1}^{2}(ka)+N_{m+1}^{2}(ka)\right]}\Biggr{)}. (36)

The brackets in the denominator can be analyzed using the integral representationBateman

Jm2(ka)+Nm2(ka)\displaystyle J^{2}_{m}(ka)+N^{2}_{m}(ka)
=8π20cosh(2mt)K0(2kasinht)𝑑t,\displaystyle=\frac{8}{\pi^{2}}\int\limits_{0}^{\infty}\cosh(2mt)K_{0}\left(2ka\sinh t\right)dt, (37)

where K0(z)K_{0}(z) is the Macdonald function. Since the product kaka is big, the integral is dominated by small t1t\ll 1, which allows to replace sinht\sinh t by tt. To perform the integration over tt it is convenient to use the following representation of the Macdonald function

K0(2kat)=1ds(s21)1/2exp(2kats).K_{0}(2kat)=\int\limits_{1}^{\infty}\frac{ds}{\left(s^{2}-1\right)^{1/2}}\exp\bigl{(}-2kats\bigr{)}. (38)

Substituting Eq. (38) into Eq. (IV.3) and integrating over tt, we get

Jm2(ka)+Nm2(ka)=4π2ka1dss(s21)1/2[s2(mka)2].J^{2}_{m}(ka)+N^{2}_{m}(ka)=\frac{4}{\pi^{2}ka}\int\limits_{1}^{\infty}\frac{dss}{\left(s^{2}-1\right)^{1/2}\left[s^{2}-\left(\frac{m}{ka}\right)^{2}\right]}. (39)

Now the evaluation of the integral is elementary and is achieved by the substitution s=(1+u2)1/2s=(1+u^{2})^{1/2}. The result reads

Jm2(ka)+Nm2(ka)=2πka(1m2k2a2)1/2.J^{2}_{m}(ka)+N^{2}_{m}(ka)=\frac{2}{\pi ka\left(1-\frac{m^{2}}{k^{2}a^{2}}\right)^{1/2}}. (40)

This result applies for m<kam<ka. For m>kam>ka the integral Eq. (39) is zero. Using the relation Eq. (40), the expression Eq. (36) for the transport cross section can be cast in the form

σtr4a=1kam=0(1m2k2a2)1/2(1(m+1)2k2a2)1/2.\frac{\sigma_{\text{tr}}}{4a}=\frac{1}{ka}\sum_{m=0}^{\infty}\left(1-\frac{m^{2}}{k^{2}a^{2}}\right)^{1/2}\left(1-\frac{(m+1)^{2}}{k^{2}a^{2}}\right)^{1/2}. (41)

If we neglect the difference between (m+1)(m+1) and mm in the second bracket, the product of the brackets will reduce to (1m2k2a2)\left(1-\frac{m^{2}}{k^{2}a^{2}}\right). Then the summation over mm will reproduce the result Eq. (33). Thus, the quantum correction to the transport cross section is due to the difference between the first and second brackets. To account for this difference we expand the second bracket as follows

(1(m+1)2k2a2)1/2\displaystyle\left(1-\frac{\left(m+1\right)^{2}}{k^{2}a^{2}}\right)^{1/2}
(1m2k2a2)1/2mk2a2(1m2k2a2)1/2.\displaystyle\approx\left(1-\frac{m^{2}}{k^{2}a^{2}}\right)^{1/2}-\frac{m}{k^{2}a^{2}\left(1-\frac{m^{2}}{k^{2}a^{2}}\right)^{1/2}}. (42)

Substituting this expansion into Eq. (41) and performing summation over mm, we arrive to the corrected expression for σtr\sigma_{\text{tr}}

σtr4a=2312ka.\frac{\sigma_{\text{tr}}}{4a}=\frac{2}{3}-\frac{1}{2ka}. (43)

We see that the correction is negative suggesting that the approach of σtr\sigma_{\text{tr}} to the limiting value is “from below”. This is the result of the expansion Eq. (42) underestimating the m=0m=0 term. Incorporating this term explicitly, we obtain

σtr4a=2312ka+4π2(ka)3[J02(ka)+N02(ka)][J12(ka)+N12(ka)].\frac{\sigma_{\text{tr}}}{4a}=\frac{2}{3}-\frac{1}{2ka}\\ +\frac{4}{\pi^{2}(ka)^{3}\left[J^{2}_{0}(ka)+N^{2}_{0}(ka)\right]\left[J_{1}^{2}(ka)+N_{1}^{2}(ka)\right]}. (44)

As seen in Fig. 4, Eq. (44) leads to the approach of σtr\sigma_{\text{tr}} to the limiting value “from above” for ka>1ka>1. Moreover, it reproduces correctly the result of numerical calculation.

Refer to caption
Figure 4: (Color online) The transport cross section in two dimensions is calculated numerically from Eq. (36). In (a), the approach to the asymptotic value of 8a/38a/3 is shown. In (b), a plot of the different contributions to the sum in Eq. (36). The blue dots represent the contribution from the terms with m1m\geq 1. The yellow dots represent the m=0m=0 term. The green dots are the sum of the blue and yellow. The red curve is a plot of Eq. (44) and is a good approximation for ka>1ka>1.

V Incorporating interactions

We will follow a transparent procedure of incorporating the interactions which was put forward in Ref. Rudin1997, for the interaction correction to the density of states and then adopted in Refs. Zala, , Gornyi2006, for the interaction correction to the conductivity. The main message of Ref. Rudin1997, , see also Ref. Yue, , is that Friedel oscillations of the electron density translate into the oscillating Hartree potential

VH(𝐫)=𝑑𝐫V(𝐫𝐫)δn(𝐫).V_{H}({\bf r})=\int d{\bf r^{\prime}}V\left({\bf r}-{\bf r^{\prime}}\right)\delta n({\bf r^{\prime}}). (45)

Here V(𝐫𝐫)V\left({\bf r}-{\bf r^{\prime}}\right) is the screened electron-electron interaction potential. Additional scattering of the incident electron from the potential VH(𝐫)V_{H}({\bf r}) modifies the scattering phases δm(m)\delta_{m}^{(m)}. In order to calculate the corresponding corrections to the phases, we employ the procedure described e.g. in Ref. Landau, .

As a first step, instead of the wave function Rm,k(0)(r)R_{m,k}^{(0)}(r), we introduce an auxiliary function χm,k(0)(r)\chi_{m,k}^{(0)}(r) defined as

Rm,k(0)(r)=χm,k(0)(r)(kr)1/2.R_{m,k}^{(0)}(r)=\frac{\chi_{m,k}^{(0)}(r)}{\left(kr\right)^{1/2}}. (46)

The function χm,k(0)\chi_{m,k}^{(0)} satisfies the equation

d2dr2χm,k(0)+[k2(m214)r2]χm,k(0)=0.\frac{d^{2}}{dr^{2}}\chi_{m,k}^{(0)}+\left[k^{2}-\frac{\left(m^{2}-\frac{1}{4}\right)}{r^{2}}\right]\chi_{m,k}^{(0)}=0. (47)

In the presence of interactions, the potential VH(r)V_{H}(r) adds to the centrifugal potential, so that Eq. (47) assumes the form

d2dr2χm,k+[k2+VH(r)(m214)r2]χm,k=0.\frac{d^{2}}{dr^{2}}\chi_{m,k}+\left[k^{2}+V_{H}(r)-\frac{\left(m^{2}-\frac{1}{4}\right)}{r^{2}}\right]\chi_{m,k}=0. (48)

Multiplying Eq. (47) by χm,k(r)\chi_{m,k}(r) and Eq. (48) by χm,k(0)(r)\chi_{m,k}^{(0)}(r) and subtracting, we arrive to the relation

ddr[dχm,kdrχm,k(0)dχm,k(0)drχm,k]=VHχm,k(0)χm,k.\frac{d}{dr}\Bigg{[}\frac{d\chi_{m,k}}{dr}\chi_{m,k}^{(0)}-\frac{d\chi_{m,k}^{(0)}}{dr}\chi_{m,k}\Bigg{]}=-V_{H}\chi_{m,k}^{(0)}\chi_{m,k}. (49)

This relation allows to find the interaction-induced correction to the scattering phase. Since this correction is small, the product χm,k(0)χm,k\chi_{m,k}^{(0)}\chi_{m,k} in the right-hand side can be replaced by (χm,k(0))2\left(\chi_{m,k}^{(0)}\right)^{2}. Then the integration of Eq. (49) from aa to \infty yields

Δδm=a𝑑rVH(r)(χm,k(0))2.\Delta\delta_{m}=-\int\limits_{a}^{\infty}dr~{}V_{H}(r)\left(\chi_{m,k}^{(0)}\right)^{2}. (50)

The correction Δδm\Delta\delta_{m} to the scattering phases give rise to the following interaction corrections to the full and to the transport scattering cross sections

δσ=4kmsin(2δm(0))Δδm,\delta\sigma=\frac{4}{k}\sum_{m}\sin\left(2\delta_{m}^{(0)}\right)\Delta\delta_{m}, (51)
δσtr=2kmsin[2(δm(0)δm+1(0))](ΔδmΔδm+1).\delta\sigma_{\text{tr}}=\frac{2}{k}\sum_{m}\sin\Biggl{[}2\left(\delta_{m}^{(0)}-\delta_{m+1}^{(0)}\right)\Biggr{]}\Bigl{(}\Delta\delta_{m}-\Delta\delta_{m+1}\Bigr{)}. (52)

Next we express the factors sin2δm(0)\sin 2\delta_{m}^{(0)} and sin2(δm(0)δm+1(0))\sin 2\left(\delta_{m}^{(0)}-\delta_{m+1}^{(0)}\right) in terms of the Bessel functions

sin2δm(0)=2Jm(ka)Nm(ka)Jm2(ka)+Nm2(ka),\sin 2\delta_{m}^{(0)}=\frac{2J_{m}(ka)N_{m}(ka)}{J_{m}^{2}(ka)+N_{m}^{2}(ka)}, (53)
sin2(δm(0)δm+1(0))\displaystyle\sin 2\left(\delta_{m}^{(0)}-\delta_{m+1}^{(0)}\right)
=Jm(ka)Nm+1(ka)Nm(ka)Jm+1(ka)Jm2(ka)+Nm2(ka)\displaystyle=\frac{J_{m}(ka)N_{m+1}(ka)-N_{m}(ka)J_{m+1}(ka)}{J_{m}^{2}(ka)+N_{m}^{2}(ka)}
×Nm(ka)Nm+1(ka)+Jm(ka)Jm+1(ka)Jm+12(ka)+Nm+12(ka).\displaystyle\times\frac{N_{m}(ka)N_{m+1}(ka)+J_{m}(ka)J_{m+1}(ka)}{J_{m+1}^{2}(ka)+N_{m+1}^{2}(ka)}. (54)

Using the fact that ka1ka\gg 1 we can simplify the expressions for δσ\delta\sigma and δσtr\delta\sigma_{\text{tr}} as follows

δσ=8cos(2ka)km=0ka(1)mΔδm,\delta\sigma=-\frac{8\cos(2ka)}{k}\sum_{m=0}^{ka}\left(-1\right)^{m}\Delta\delta_{m}, (55)
δσtr=8km=0kamka[1(mka)2]1/2(ΔδmΔδm+1).\delta\sigma_{\text{tr}}=\frac{8}{k}\sum_{m=0}^{ka}\frac{m}{ka}\Biggl{[}1-\left(\frac{m}{ka}\right)^{2}\Biggr{]}^{1/2}\left(\Delta\delta_{m}-\Delta\delta_{m+1}\right). (56)

We see that the two expressions are very different. Since Δδm\Delta\delta_{m} is a smooth function of mm, the terms in Eq. (55) cancel out. The same smoothness of Δδm\Delta\delta_{m} allows to replace the sum over mm in Eq. (56) by the integral

δσtr=8k0ka𝑑m(mka)[1(mka)2]1/2Δδmm.\delta\sigma_{\text{tr}}=-\frac{8}{k}\int\limits_{0}^{ka}dm\left(\frac{m}{ka}\right)\Biggl{[}1-\left(\frac{m}{ka}\right)^{2}\Biggr{]}^{1/2}\frac{\partial\Delta\delta_{m}}{\partial m}. (57)

Next we argue that relevant values of mm are much smaller than kaka. This allows to replace the square bracket by 11 and transform Eq. (57) by parts. This yields

δσtr=8k2a0𝑑m(Δδm).\delta\sigma_{\text{tr}}=\frac{8}{k^{2}a}\int\limits_{0}^{\infty}dm\left(\Delta\delta_{m}\right). (58)

To analyze the dependence of Δδm\Delta\delta_{m} given by Eq. (50) on the wave vector, kk, of the incident electron we recall that the potential, VH(r)V_{H}(r), is proportional to electron density given by Eq. (4). To pinpoint the origin of the anomaly at k=kFk=k_{\scriptscriptstyle F} it is more convenient to study the derivative ΔδmkF\frac{\partial\Delta\delta_{m}}{\partial k_{\scriptscriptstyle F}}. Within a factor, this derivative is given by

ΔδmkFa𝑑r[rRm,k2(r)]mRm,kF2(r).\frac{\partial\Delta\delta_{m}}{\partial k_{\scriptscriptstyle F}}\propto\int\limits_{a}^{\infty}dr\Bigl{[}rR_{m,k}^{2}(r)\Bigr{]}\sum_{m^{\prime}}R_{m^{\prime},k_{\scriptscriptstyle F}}^{2}(r). (59)

From the semiclassical form of the radial wave function Eq. (5) we conclude that the product Rm,k2(r)Rm,kF2(r)R_{m,k}^{2}(r)R_{m^{\prime},k_{\scriptscriptstyle F}}^{2}(r) contains a slow part

cos2{ar𝑑r[(k2m2r2)1/2(kF2m2r2)1/2]}(k2r2m2r2)1/2(kF2r2m2)1/2.\frac{\cos 2\Bigl{\{}\int\limits_{a}^{r}dr^{\prime}\Bigl{[}\left(k^{2}-\frac{m^{2}}{r^{2}}\right)^{1/2}-\left(k_{\scriptscriptstyle F}^{2}-\frac{m^{\prime 2}}{r^{2}}\right)^{1/2}\Bigr{]}\Bigr{\}}}{\left(k^{2}r^{2}-\frac{m^{2}}{r^{2}}\right)^{1/2}\left(k_{\scriptscriptstyle F}^{2}r^{2}-m^{\prime 2}\right)^{1/2}}. (60)

Since we assumed that mm and mm^{\prime} are both much smaller than kaka, the above expression can be simplified as follows

cos2[(kkF)(ra)m2m22k(1a1r)]kF2r2.\frac{\cos 2\Bigl{[}(k-k_{\scriptscriptstyle F})(r-a)-\frac{m^{2}-m^{\prime 2}}{2k}\Bigl{(}\frac{1}{a}-\frac{1}{r}\Bigr{)}\Bigr{]}}{k_{\scriptscriptstyle F}^{2}r^{2}}. (61)

It is natural to measure the radial coordinate, rr, from r=ar=a. Combining Eqs. (58), (59), and (61), we arrive to the following expression for the derivative of δσtr\delta\sigma_{\text{tr}} with respect to kFk_{\scriptscriptstyle F}

δσtrkF0𝑑m0𝑑m0dρρ+acos2[(kkF)m2m22ka(ρ+a)]ρ.\frac{\partial\delta\sigma_{\text{tr}}}{\partial k_{\scriptscriptstyle F}}\propto\mkern-6.0mu\int\limits_{0}^{\infty}\mkern-6.0mudm\mkern-6.0mu\int\limits_{0}^{\infty}\mkern-6.0mudm^{\prime}\mkern-6.0mu\int\limits_{0}^{\infty}\frac{d\rho}{\rho+a}\cos 2\Bigl{[}(k-k_{\scriptscriptstyle F})-\frac{m^{2}-m^{\prime 2}}{2ka(\rho+a)}\Bigr{]}\rho. (62)

We can now perform the integration over mm and mm^{\prime} explicitly. This yields

δσtrkF2πka0dρρcos2[(kkF)ρ].\frac{\partial\delta\sigma_{\text{tr}}}{\partial k_{\scriptscriptstyle F}}\propto 2\pi ka\int\limits_{0}^{\infty}\frac{d\rho}{\rho}\cos 2\Big{[}\left(k-k_{\scriptscriptstyle F}\right)\rho\Big{]}. (63)

Since the expression for the Friedel oscillations applies for ρ=(ra)kF1\rho=(r-a)\gg k_{\scriptscriptstyle F}^{-1}, the lower limit in the integral should be chosen to be ρkF1\rho\sim k_{\scriptscriptstyle F}^{-1}. Thus, we arrive to the final result

δσtrkFln(kFkkF)=ln(EFε).\frac{\partial\delta\sigma_{\text{tr}}}{\partial k_{\scriptscriptstyle F}}\propto\ln\Bigl{(}\frac{k_{\scriptscriptstyle F}}{k-k_{\scriptscriptstyle F}}\Bigr{)}=\ln\Bigl{(}\frac{E_{\scriptscriptstyle F}}{\varepsilon}\Bigr{)}. (64)

From Eq. (64) we conclude that the interaction correction to the transport cross section has the form δσtrεln(EFε)\delta\sigma_{\text{tr}}\propto\varepsilon\ln\Bigl{(}\frac{E_{\scriptscriptstyle F}}{\varepsilon}\Bigr{)}. Recall, that for point-like impurityZala the interaction correction has the form δσtrε\delta\sigma_{\text{tr}}\propto\varepsilon. Thus, the enhancement of the interaction correction in the case of scattering from the disk amounts to the logarithmic factor.

Refer to caption
Figure 5: (Color online) Full cross section (blue curve) and scattering cross section (yellow curve) are plotted in log-log scale from Eqs. (13) and (15), respectively. With increasing the dimensionless parameter, kaka, the transport cross section approaches the limiting value, 8a3\frac{8a}{3}, faster than the full cross section approaches the limiting value, 4a4a.

VI Concluding remarks

(i) In Fig. 5 the numerical results for the full and transport scattering cross sections are presented in log-log scale. It is apparent that σ\sigma and σtr\sigma_{\text{tr}} approach their limiting values σ=4a\sigma=4a and σtr=83a\sigma_{\text{tr}}=\frac{8}{3}a, respectively, at very different rates. Naturally, at small ka1ka\ll 1 both cross sections coincide. However, at large kaka, the transport cross section saturates much faster than the full cross section. On the contrary, we have demonstrated that while the interaction correction to σtr\sigma_{\text{tr}} is a singular function of kkFk-k_{\scriptscriptstyle F}, the interaction correction to σ\sigma vanishes.

(ii) Integration over momenta mm and mm^{\prime} in Eq. (62) leads to the slow energy dependence of the correction δσtrkF\frac{\partial\delta\sigma_{\text{tr}}}{\partial k_{\scriptscriptstyle F}}. Note, that this integration misses a contribution from the terms m=mm=m^{\prime}, which is the consequence of the discreteness of the angular momentum. This contribution is comparable to the main contribution Eq. (64) for the following reason. In integration, the characteristic values of mm and mm^{\prime} are (ka)1/2\sim(ka)^{1/2}, while the number of terms with m=mm=m^{\prime} is ka\sim ka. On the other hand, the contribution of these terms possesses a sharp energy dependence. This contribution can be cast in the form

δσ~trkF0dzcoszz+2(kkF)a=0dzsinz[z+2(kkF)a]2.\frac{\partial\delta{\tilde{\sigma}}_{\text{tr}}}{\partial k_{\scriptscriptstyle F}}\propto\int\limits_{0}^{\infty}\frac{dz\cos z}{z+2\left(k-k_{\scriptscriptstyle F}\right)a}=\int\limits_{0}^{\infty}\frac{dz\sin z}{\Bigl{[}z+2\left(k-k_{\scriptscriptstyle F}\right)a\Bigr{]}^{2}}. (65)

Eq. (65) yields a narrow peak of a width (kkF)a1\left(k-k_{\scriptscriptstyle F}\right)\sim a^{-1} and the corresponding energy scale εEFka\varepsilon\sim\frac{E_{\scriptscriptstyle F}}{ka}. At large (kkF)a1\left(k-k_{\scriptscriptstyle F}\right)a\gg 1 the correction Eq. (65) falls off as 14(kkF)2a2\frac{1}{4\left(k-k_{\scriptscriptstyle F}\right)^{2}a^{2}}. Note also, that the correction Eq. (65) originates from the specific behavior of the Friedel oscillations Eq. (9) from the disk as compared to the Friedel oscillations from the point impurity. It thus contains the disk radius, aa. Upon integrating Eq. (65), we get the following shape of the peak in the transport cross section

δσ~tr0dzsinzz+2(kkF)a.\delta{\tilde{\sigma}}_{\text{tr}}\propto\int\limits_{0}^{\infty}\frac{dz\sin z}{z+2\left(k-k_{\scriptscriptstyle F}\right)a}. (66)

The missing prefactor in Eq. (66) contains the product of the electron-electron interaction constantZala and the density of states, ν0\nu_{0}. The first factor originates from the proportionality between VH(r)V_{H}(r) and δn(r)\delta n(r) in Eq. (50), while the second factor comes from the sum in Eq. (59), which emerges upon taking the derivative with respect to kFk_{\scriptscriptstyle F}.

(iii) We emphasize the difference between the interaction corrections for the cases of a point-like impurity and of a big disk with ka1ka\gg 1. For a point-like impurity,Zala the corrections to the full and to the transport scattering cross sections are related as δσtr=2δσ\delta\sigma_{\text{tr}}=2\delta\sigma. This can also be seen from Eq. (50). For ka1ka\ll 1, the correction, Δδ0\Delta\delta_{0}, is much bigger than Δδm\Delta\delta_{m} for m0m\neq 0. By contrast, for ka1ka\gg 1, the correction, δσ\delta\sigma, is small in parameter 1ka\frac{1}{ka}.

VII Acknowledgements

N.F. acknowledges the support of the National Science Foundation (NSF) award No. 1950409. M.E.R. was supported by the Department of Energy, Office of Basic Energy Sciences, Grant No. DE-FG02-06ER46313.

References

  • (1) F. Stern “Polarizability of a Two-Dimensional Electron Gas,” Phys. Rev. Lett. 18, 546 (1967).
  • (2) A. V. Chubukov and D. L. Maslov, “Nonanalytic corrections to the Fermi-liquid behavior,” Phys. Rev. B 68, 155113 (2003).
  • (3) G. W. Martin, D. L. Maslov, and M. Yu. Reizer, “Quantum magneto-oscillations in a two-dimensional Fermi liquid,” Phys. Rev. B 68, 241309(R) (2003).
  • (4) A. Gold and V. T. Dolgopolov, “Temperature dependence of the conductivity for the two-dimensional electron gas: Analytical results for low temperatures,” Phys. Rev. B 33 1076 (1986).
  • (5) A. M. Rudin, I. L. Aleiner, and L. I. Glazman, “Tunneling zero-bias anomaly in the quasiballistic regime,” Phys. Rev. B 55, 9322 (1997).
  • (6) G. Zala, B. N. Narozhny, and I. L. Aleiner, “Interaction corrections at intermediate temperatures: Longitudinal conductivity and kinetic equation,” Phys. Rev. B 64, 214204 (2001).
  • (7) Y. Adamov, I. V. Gornyi, and A. D. Mirlin, “Interaction effects on magneto-oscillations in a two-dimensional electron gas,” Phys. Rev. B 73, 045426 (2006).
  • (8) L. D. Landau and E. M. Lifshitz, Quantum Mechanics (Non-Relativistic Theory), Course of Theoretical Physics Vol. 3, 3rd ed. (Pergamon, 1985).
  • (9) I. R. Lapidus, “Scattering by two-dimensional circular barrier, hard circle, and delta function ring potentials,” Am. J. Phys. 54, 459 (1986).
  • (10) K. A. Matveev, Dongxiao Yue, and L. I. Glazman, “Tunneling in one-dimensional non-Luttinger electron liquid,” Phys. Rev. Lett. 71, 3351 (1993).
  • (11) H. Bateman, A. Erdélyi, W. Magnus, F. Oberhettinger, and F. G. Tricomi, “Higher transcendental functions,” New York: McGraw-Hill. 2 (1955).