• Aucun résultat trouvé

Effect of interactions on the admittance of ballistic wires

N/A
N/A
Protected

Academic year: 2022

Partager "Effect of interactions on the admittance of ballistic wires"

Copied!
5
0
0

Texte intégral

(1)

Article

Reference

Effect of interactions on the admittance of ballistic wires

BLANTER, Iaroslav, BUTTIKER, Markus

Abstract

A self-consistent theory of the admittance of a perfect ballistic, locally charge neutral wire is proposed. Compared to a non-interacting theory, screening effects drastically change the frequency behavior of the conductance. In the single-channel case the frequency dependence of the admittance is monotonic, while for two or more channels collective interchannel excitations lead to resonant structures in the admittance. The imaginary part of the admittance is typically positive, but can become negative near resonances.

BLANTER, Iaroslav, BUTTIKER, Markus. Effect of interactions on the admittance of ballistic wires. Europhysics Letters , 1998, vol. 42, no. 5, p. 535-540

DOI : 10.1209/epl/i1998-00284-x arxiv : cond-mat/9706070

Available at:

http://archive-ouverte.unige.ch/unige:4218

Disclaimer: layout of this document may differ from the published version.

(2)

arXiv:cond-mat/9706070v2 [cond-mat.mes-hall] 23 Jun 1998

Effect of Interactions on the Admittance of Ballistic Wires

Ya. M. Blanter and M. B¨uttiker

D´epartement de Physique Th´eorique, Universit´e de Gen`eve, CH-1211, Gen`eve 4, Switzerland.

(February 1, 2008)

A self-consistent theory of the admittance of a perfect ballistic, locally charge neutral wire is proposed. Compared to a non-interacting theory, screening effects drastically change the frequency behavior of the conductance. In the single-channel case the frequency dependence of the admittance is monotonic, while for two or more channels collective interchannel excitations lead to resonant structures in the admittance. The imaginary part of the admittance is typically positive, but can become negative near resonances.

PACS numbers: 73.23.Ad, 73.61.-r

The ac conductance (admittance) in mesoscopic sys- tems attracted recently strong interest, mostly due to the finite-frequency measurements of Aharonov-Bohm oscil- lations in rings [1] and noise in diffusive metallic wires [2]. The theoretical investigation of the problem raises an important question: Standard calculations of the con- ductance, employing either a scattering (Landauer) or a linear response (Kubo) approach, describe the current in response to theexternalelectric field, assuming the latter to be uniform (linear response) or ignoring the actual dis- tribution of the potential (dc scattering approach). How- ever, in realistic systems the potential isnotuniform due to the screening effects. In dc transport the actual poten- tial distribution is unimportant for the evaluation of the conductance due to the Einstein relation. In contrast, the ac response is strongly sensitive to the distribution of the potential inside the sample [3,4]. In its turn, the potential profile is related via the Poisson equation to the electron density. Thus, the admittance has to be found self-consistently [5].

Indeed, early attempts to generalize dc conductance calculations to the ac case (see e.g.[6]) within the free- electron approach proved to be not self-consistent and failed to conserve current. The sensitivity of the ac conductance to different electric field configurations is illustrated in Ref. [7]. The construction of a current- conserving theory is not an easy task. Presently, two approaches are available in the literature. First, the self-consistent ac scattering approach [3,4] allowed to study the low-frequency admittance and to express it through the scattering matrix for an arbitrary system [4,8]. For arbitrary frequencies, the admittance is ex- pressed through functional derivatives of the scattering matrix with respect to the local electric potential. Alter- natively, one can employ the methods of non-equilibrium statistical mechanics, and express the ac response of an interacting system through the Green’s functions in a Keldysh formalism [9,10]. Although these methods are quite powerful, their application to specific prob- lems meets substantial technical difficulties. The high frequency response of coupled infinitely extended one- channel liquids has been investigated by Matveev and Glazman [11]. Here we are particularly interested in

wires connected to reservoirs. Compared to infinitely extended wires the presence of reservoirs modifies the low frequency response, possibly up to frequencies deter- mined by a transit time.

Below we develop a self-consistent theory for the ad- mittance of a perfect ballistic wire (directed along the x-axis) of a length L and a cross-section S ≪ L2 (3D) or width W ≪ L (2D), placed between two reservoirs [12,13]. The wire is assumed to be shorter than all the lengths associated with inelastic scattering. Here we treat interactions in the random phase approximation (RPA). More specifically, we are interested in the limit where the wire is locally charge neutral. This is realistic in two situations: (i) Suppose for a moment that the wire is coupled to a back-gate through a capacitancecper unit length. In the one-channel case, this system is equivalent to the Luttinger model with short-range interactions [14];

the interaction constantgand the capacitance are related byg2 = (1 +e2/(πvFc))−1. Here vF is the Fermi veloc- ity which determines the density of states 1/πvF. The charge neutral case corresponds to the zero capacitance limit (org= 0 limit), that is, the back gate is at a very large distance from the wire. This example also shows that the condition of applicability of RPA e2 ≪ vF is compatible with a small interaction constantg since the capacitance can be very small. Based on this model the results presented below are valid for frequencies up to vF/R, where R ≪ L is the distance between the wire and the gate. The case of many channels can not be re- duced to the Luttinger model. (ii) In the absence of a back-gate (or forR≫L), the results presented below are valid for frequencies below the plasma modes of a wire.

For one channel these frequencies are of orderωp∼vF/L (seee.g.[11,15]); for a multi-channel wire the highest ly- ing plasmon branch (vF/L) provides the frequency which limits the results presented below.

We show that screening plays a crucial role for the fre- quency dependence of the admittance: a wire with only one transverse channel exhibits a monotonic frequency dependence, while a wire with several transverse chan- nels shows resonant structures in the admittance, due to collective interchannel excitations in which an accumula- tion of charge in one channel is locally compensated by a

(3)

charge depletion in another channel. As a consequence, the imaginary part of the admittance changes sign as a function of frequency.

General formulation. The ac transport in a 1D per- fect wire is determined by the following system of equa- tions for the local current j, the particle densityρ, and the electric potentialϕ/e:

j =jp−(4πe)−1∇∂tϕ; jp(x) =eX

a

va ρ+a −ρa

; (1)

∆ϕ=−4πe2X

a

ρ+aa

; (2)

−∂xjp=e∂t

X

a

ρ+aa

; (3)

ρ±a(x, t) =ρ±0a(x, t)

− Z L

0

Π±a(x, x, t−t)ϕ(x, y= 0, z= 0, t)dxdt. (4) The index a labels transverse channels (1 ≤ a ≤ N), and ρ±a denotes the excess density (with respect to the positive background) of right/left-movers in the channel a, which depends only on the coordinatex. The veloci- ties va of left- and right-movers in the same channel co- incide, va = (πSνa)−1. Here νa is the density of states in the channela; the total density of states is given by ν=Pνa. Eqs. (1) - (4) are valid within the ballistic wire extending fromx= 0 tox=L. Furthermore, the quan- tities ρ±0a are “bare” densities of particles in the channel a injected from the left/right reservoir; the distribution function of these particles is determined by the distri- bution function of the left/right reservoir at the time of injection. Specifying the chemical potentials in the left/right reservoirsµL=V(t),µR= 0, we obtain

ρ±0a =1 2νa

Z

0

fL,R(ǫ)dǫ,

and it follows ρ+0a(x, t) =νaV(t−x/va)/2 andρ0a = 0.

Note that for x = L the bare density of right-movers does not vanish: We do not consider in detail the tran- sition region between the wire and reservoirs, where the electrons are distributed over many quantum channels.

Finally, Π±a(x, x, t−t) is the polarization function, re- sponsible for the density of right/left-movers induced in the channelaat the pointxand timetdue to a potential perturbationϕatx, t. If spatial and temporal structure on the scale ofp−1F (Friedel oscillations) andǫ−1F , respec- tively, can be neglected, the polarization function quite generally has the form

Π±a(x, x, t) = (νa/2)

δ(x−x)δ(t)−∂tPa±(x, x, t) , (5) where the function Pa± is the conditional probability to find a right/left-moving particle in the channelaatxat timet, if it was atx att = 0. For a ballistic channel we have obviously

Pa±(x, x, t) =θ(t)δ(x±vat−x). (6) Now we return to our system of equations. First, we note that the Poisson equation (2) and the continuity equation (3) together with the definition of the current (1) imply∂j/∂x= 0, i.e. the current is conserved and does not depend on the space point. The particle current jp is generally not conserved, and the displacement cur- rent is required for the self-consistent treatment. How- ever, if the system is locally charge neutral, the case of interest here, the term ∆ϕin Eq.(2) vanishes simultane- ously with the displacement current in Eq. (1). Further- more, we note that the system of equations (1)-(4) is ex- cessive: 2N+ 2 equations (2), (3), (4) contain 2N+ 1 unknown fieldsϕ and ρ±a. Eq. (1) is already a conse- quence of the continuity equation and the Poisson equa- tion. It is thus sufficient to consider Eqs. (1), (2) and (4): The fact that the resulting current is position inde- pendent can be used as simple test of consistency.

It is convenient to Fourier-transform the equations with respect to time, and to introduce new variables:

the full density in the channel a, ρa = ρ+aa, and the density difference ζa = ρ+a −ρa. The bare density injected by the reservoir into channel a in response to a potential oscillation Vω in that reservoir is ρ0a(x) = (νaVω/2) exp(iωx/va). The combined contribution of the probabilitiesP± gives after Fourier transform a term

pa(x) = (iωνa/2va) exp (iω|x|/va)

in the polarization function. It is useful also to introduce the operator ˆQa,

ag=νag(x) + Z L

0

pa(x−x)g(x)dx. With these abbreviations we obtain

j=eX

a

vaζa(x); ζa(x) =ρ0a(x)

− Z L

0

sign(x−x)pa(x−x)ϕ(x)dx; (7) X

a

ρa(x) = 0; (8)

ρa(x) =ρ0a(x)−Qˆaϕ(x), (9) whereϕ(x) ≡ϕ(x, y = 0, z = 0). Combining equations (8) and (9) gives a closed integral equation for the po- tential, generated in response to the injected density,

X

a

aϕ=X

a

ρ0a. (10)

Now we have to solve the equation (10) for the po- tential, and then calculate the current from Eq. (7).

The admittance is determined by G(ω) = ejS/Vω. We

(4)

re-emphasize again that the current and thus the admit- tance do not depend on the space point x, which is a check of the consistency of our approach.

One channel. For the case of one channel with the velocityv the solution of Eq. (10) has the form

ϕ(x) =Vω

1− 2v

iωL −1

1−x L − v

iωL

. (11) For ω ≪v/L the potential is screened, and everywhere in the wire is close to one half of the external voltage. On the other hand forω≫v/Lthe potential drops linearly.

Eq. (10) describes thus the crossover from a uniform po- tential at low frequencies to a uniform electric field at high frequencies. Calculating the admittance, we obtain

G(ω) = e2

1−iωL 2v

−1

. (12)

The imaginary part of the admittance is positive (in- ductive). For low frequencies we reproduce the behavior found previously in Ref. [8],G(ω) =e2N/π−iωE, with an emittance

E=−e2

4νLS. (13)

Two channels. The Poisson equation for the poten- tial (10) for a wire with two channels a and b can be solved by noticing that

2+v2a2/∂x2) ˆQaϕ=νava2ϕ′′. This implies that the potential has the form

ϕ(x) =α+βx+γexp(iξx) +δexp(−iξx), where the coefficientsα,β,γandδare determined by the substitution of this ansatz into Eq. (10). Interestingly, now, in addition to the constant and linear part, the po- tential acquires also a spatially oscillating part with a wave vector given by ξ = ω/(vavb)1/2. The oscillatory structure of the potential is also manifest in the admit- tance. Indeed, we find a non-trivial dependence of the conductance on the wave vector ξ and the density of states ratioη= (va/vb)1/2,

G(ω) = e2

2π(η2+ 1)

−iξLη

2 +(η3+ 1)(η+ 1)−(η3−1)(η−1) exp(iξL) (η+ 1)2−(η−1)2exp(iξL)

−1

(14)

The conductance is symmetric with respect to the re- placementη→η−1, as it must be. For low frequencies we reproduce again the static quantized conductance e2/π and the emittance (13). The real part is strictly posi- tive, although now both real and imaginary part exhibit oscillations on top of the monotonic behavior found for the single channel wire. In the limit va = vb, which corresponds to a spin degenerate one-channel conductor, these oscillations vanish. The behavior of the admit- tance as a function of frequency for different values of the parameterη is shown in Fig. 1. Note that the imag- inary part may change sign in the vicinity of the points ω∼2πn(vavb)1/2/L(see below), ifη is large enough. A largeηoccurs for Fermi energies just above the threshold of the second conductance channel.

The oscillatory structure of the admittance can be un- derstood by investigating the poles of the admittance. In the limiting case η ≪1, we obtain for the spectrum of collective modes,

ω= (vavb)1/2 L

2πn(1−η2)−2iη

, n∈Z.

The purely imaginary eigenvalue with n = 0, as for one channel, corresponds to charge relaxation between the wire and the reservoirs via ballistic motion, while the resonances for n 6= 0 correspond to nearly neu- tral interchannel excitations. These modes are essen- tially standing waves induced in both channels simulta- neously but with densities of opposite signs,ρa =−ρb

exp(iξx)−exp(iξ(L−x)). The decay of these excita- tions, Im ω =−2va/L, is determined by the carriers in the channel with the lower velocity. Interestingly, we find that due to the coupling to reservoirs all the collec- tive modes are damped with a relaxation constant which is the larger the shorter the wire is.

Many channels. ForN >1 channels the potential has the form

ϕ(x) =α+βx+

N−1

X

i=1

iexp(iωλix) +δiexp(−iωλix)], and the positive quantitiesλi are solutions of the equa- tion

X

a

va

1−v2aλ2 = 0. (15) For arbitrary N and ω, further analytical progress is hard, but the problem can be solved for low frequencies ω≪min(va/L). We obtain

G(ω) = e2N

2π +iωe2

4νSL−ω2e2π

8 (SL)2X

a

νa2+. . . . (16) In conclusion, we investigated the admittance of a per- fect ballistic wire in the frequency range belowvF/L. We showed that the screening effects are very important for the admittance, and provide the current conservation.

(5)

For one channel, the admittance is a monotonic func- tion of frequency, whereas for two or more channels it contains also oscillatory components due to the density redistribution between different channels. In particular, the imaginary part of the admittance is generally positive (inductive-like), but can change sign and become capaci- tive in the vicinity of resonances due to the interchannel excitations. The wire exhibits resonances due to damped collective modes. The resonant effects predicted can be measured experimentally; the collective mode frequen- cies depend on the relative electron concentration in the different channels and near the threshold for a new quan- tum channel, where the resonances are most pronounced, can be made very small.

We thank L. I. Glazman, D. E. Khmel’nitskii, and K. A. Matveev for useful discussions. The work was sup- ported by the Swiss National Science Foundation.

Re G

5 10 15 20

0.5 1.

1.5 2.

a)

1 2

3

Im G

5 10 15 20

-0.25 0.25 0.5 0.75 1.

b)

1

2

3

FIG. 1. Real (a) and imaginary (b) parts of the conduc- tance in units of e2/2π for the two-channel case as a func- tion of the parameter Ω = ωL/(vavb)1/2; the parameter η = (va/vb)1/2 is equal to 1 (Curve 1), 5 (2) and 20 (3).

Resonances are due to the interchannel excitations.

[1] J. B. Pieper and J. C. Price, Phys. Rev. Lett.72, 3586 (1994).

[2] R. J. Schoelkopf, P. J. Burke, A. A. Kozhevnikov, D. E. Prober, and M. J. Rooks, Phys. Rev. Lett. 78, 3370 (1997).

[3] M. B¨uttiker, H. Thomas, and A. Prˆetre, Phys. Lett. A 180, 364 (1993); T. Christen and M. B¨uttiker, Phys. Rev.

Lett.77, 143 (1996); P. W. Brouwer and M. B¨uttiker, Eu- rophys. Lett.37, 441 (1997); J. Wang and H. Guo, Phys.

Rev. B54, R11090 (1996).

[4] M. B¨uttiker, J. Phys. Cond. Matter5, 9361 (1993).

[5] Many questions, like the contactless ac response, deter- mined by the plasmon dispersion, in quantized Hall sys- tems, cannot be addressed without a self-consistent treat- ment: see e.g. D. V. Mast, A. J. Dahm, and A. L. Fet- ter, Phys. Rev. Lett.54, 1706 (1985); V. A. Volkov and S. A. Mikhailov, Zh. Eksp. Teor. Fiz. 94, 217 (1988) [Sov. Phys. JETP67, 1639 (1988)]; Ya. M. Blanter and Yu. E. Lozovik, Physica B182, 254 (1992); U. Z¨ulicke and A. H. MacDonald, Phys. Rev. B54, 16813 (1996).

[6] B. Velick´y, J. Maˇsek and B. Kramer, Phys. Lett. A140, 447 (1989); H. C. Liu, Phys. Rev. B 43, 12538 (1991);

Y. Fu and S. F. Dudley, Phys. Rev. Lett.70, 65 (1993);

L. Y. Chen and S. C. Ying, J. Phys. Cond. Matter 6, 5061 (1994).

[7] V. A. Sablikov and E. V. Chenskii, Pis’ma Zh. Eksp.

Teor. Fiz.60, 397 (1994) [JETP Lett.60, 410 (1994)].

[8] M. B¨uttiker and T. Christen, in Quantum Transport in Semiconductor Submicron Structures, Ed. by B. Kramer (Kluwer, Dordrecht, 1996), p.263.

[9] N. Wingreen, A.-P. Jauho, and Y. Meir, Phys. Rev. B 48, 8487 (1993).

[10] C. Bruder and H. Schoeller, Phys. Rev. Lett. 72, 1076 (1994).

[11] K. A. Matveev and L. I. Glazman, Physica B189, 266 (1993).

[12] In the following we use a notation appropriate for 3D; in a 2D geometry S should be replaced by W. The num- ber of transverse channels is of orderN∼p2FS (3D) or N∼pFW (2D). We consider the case of zero tempera- ture, and use the system of units with ¯h=kB= 1.

[13] This treatment resembles one performed for low fre- quency ac transport in Ref. [4] and for dc transport, Y. B. Levinson, Zh. Eksp. Teor. Fiz. 95, 2175 (1989) [Sov. Phys. – JETP 68, 1257 (1989)]; A. Kawabata, J.

Phys. Soc. Jap. 65, 30 (1996); cond-mat/9701171 (un- published).

[14] Ya. M. Blanter, F. W. J. Hekking, and M. B¨uttiker, cond- mat/9710299 (unpublished).

[15] A. V. Klyuchnik, Yu. E. Lozovik, and A. B. Oparin, Phys.

Lett. A179, 372 (1993).

Références

Documents relatifs

Comparison between experimental and FE model results calculated from measured and manufacturer- provided loss factors: (a) conductance, (b) susceptance... 4.3 Comparison

Electrical transport in ultrathin Metal-insulator-semiconductor (MIS) tunnel junctions is analyzed using the temperature dependence of current density and admittance characteristics,

Finally, the gradient calculation is used in a classical optimization method in order to design the admittance distribution of a low-height tramway noise barrier of di

Based on complex space discretized hysteresis model of the ferromagnetic materials (including both macroscopic and microscopic eddy currents contributions), a

Ea, the clouds with the largest anisotropy energy, Ea max ' are blocked at the highest temperatures, this giving rise to the susceptibility maximum which "defines" T

Starting from the available Kramers-Kronig relations for natural optical activity and the rigorously determined static value of the chirality pseudoscalar (in the Tellegen

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des