Coulomb drag of Luttinger liquids and quantum-Hall edges Karsten Flensberg arXiv:cond-mat/9802220v2 [cond-mat.mes-hall] 2 Jun 1998 Danish Institute of Fundamental Metrology, Building 307, Anker Engelunds Vej 1, DK-2800 Lyngby, Denmark (February 19, 1998) We study the transconductance for two coupled one-dimensional wires or edge states described by Luttinger liquid models. The wires are assumed to interact over a finite segment. We find for the interaction parameter g = 1/2 that the drag rate is finite at zero temperature, which cannot occur in a Fermi-liquid system. The zero temperature drag is, however, cut off at low temperature due to the finite length of the wires. We also consider edge states in the fractional quantum Hall regime, and we suggest that the low temperature enhancement of the drag effect might be seen in the fractional quantum Hall regime. in the linear and the non-linear regimes in Ref. [13]. In Ref. [14] the non-linear transresistance for two LLs at zero temperature was studied and it was argued that absolute drag (equal currents in the two wires) is possible. Here we show a related effect for the linear conductance. One-dimensional (1D) systems have attracted much attention since the advances in lithographical fabrication techniques have made it possible to study systems such as e.g. quantum point contacts, quantum wires, quantum dots, and nano-tubes. Interacting 1D systems are particularly interesting, because they are believed to be Luttinger liquids (LLs) and thus exhibit non-Fermi liquid behavior. However, it is still not clear to what extent the non-Fermi liquid behavior can be seen in a transport experiment for clean systems, i.e. without impurity scattering. Several authors [1] have shown that the interactions do not influence the conductance, and the reason for a possible interaction induced effect observed [2] at finite temperature remains unexplained. We concentrate on the temperature dependence of the linear response. Utilizing a mapping to two decoupled LLs, it is shown that a LL description shows zero temperature drag for the case where the coupling parameter g = 1/2 [15]. Moreover, we find an interesting dependence of the length of the interaction region and a regime for g < 1/2, where the drag current is almost equal to the drive current. Our results are applied to edge states in coupled FQH systems with a narrow contriction, which is different from the situation of the Ref. [11]. Another very interesting testing ground for LL behavior is edge states in the fractional quantum Hall (FQH) regime, where edge states can be described as chiral LLs [3]. Surprisingly, experiments have shown that the tunneling between two edge states follows the LL behavior [4] both for compressible and incompressible states. Theoretical descriptions in terms of 1D edge channels have been developed [5], and several authors [6,7] have recently extended this idea and calculated tunneling density of states which offers an explanation of the observed characteristics. LI LW FIG. 1. Schematic outline of the geometry considered here. The two wires are of length LW , but they interact only in a region of length LI . In this paper, another experiment which measures the interaction effects is suggested, namely a Coulomb drag experiment. Coulomb drag has during recent years proved to be a powerful tool for studying interaction and screening properties of coupled two-dimensional electron systems [8], including phonon-mediated interactions [9], coupled QH systems both in the integer regime [10] and in the fractional regime where recent experiments show Coulomb drag that saturates at the lowest temperatures at filling factor close to one half [11]. This type of behavior cannot be explained within the present weak coupling theories for bulk composite-Fermion drag [12]. In our model the two spin-less LLs are coupled by Coulomb interactions, and interwire tunneling is neglected. They are coupled in a finite region of length LI . The wires have lengths LW and it is assumed that the intrawire interaction is constant in this region, see Fig. 1. Thus the results are valid for temperatures or voltages larger than the cut-off given by the energy of charge excitations with wavelengths of order LW , i.e., T > TW = hvF /LW , where vF is a Fermi velocity. ¯ The Hamiltonians for the separate systems are those of two LLs (i = 1, 2) (we use h = kB = 1) ¯ 1D drag between two Fermi liquids has been studied 1 H0i = vi 2 dx [Pi (x)]2 + 1 2 , 2 [∂x φi (x)] gi can be neglected. The final expression for the interaction now reads (1) where gi , vi are the interaction parameters and Fermi velocities, respectively, and where the densities, ρL and ρR , of left and right movers respectively, enter as √ ∂x φi (x) = [ρLi (x) + ρRi (x)] π and P (x) = −[ρRi (x) − √ ρLi (x)] π. The fields φi and Pi form conjugate variables: [φi (x), Pj (x′ )] = iδij δ(x − x′ ). The Coulomb interaction between the two wires is given by Hint = dxdy U12 (x, y)ρ1 (x)ρ2 (y). Hint = 1 dxdy U12 (x, y) × 2π 2 α2 √ cos[2(kF 1 x − kF 2 y) + 2 π(φ1 (0) − φ2 (0))] √ + cos[2(kF 1 x + kF 2 y) + 2 π(φ1 (0) + φ2 (0))] . (6) Now we transform the Hamiltonian to that of two (interacting) LLs scattering against a single impurity potential. Define the new field operators Φ = φ1 + φ2 , Note that the interwire interaction U12 (x, y) only acts within a region of length LI . Here the subsystem densities are given by P = (P1 + P2 )/2, (7a) Θ = φ1 − φ2 , (2) Π = (P1 − P2 )/2, (7b) defined such that Φ, P and Θ, Π are conjugate pairs. The interwire interaction term becomes particularly simple in this basis and the Hamiltonian transforms to ρi (x) = ρLi (x) + ρRi (x) + Ψ† (x)ΨLi (x) + Ψ† (x)ΨRi (x), Li Ri H = H0 + H ′ + Hint , (3) where where ΨL(R)i are Fermion operators corresponding to the left(right) movers. In the bosonization language these are given by [16] (one set for each wire)ΨL(R) (x) = √ x exp ±ikF x ∓ 2π −∞ dyρL(R) (y) / 2πα, α is a cut-off parameter [17]. We separate the interwire interaction in 4 terms describing the different possibilities of forward and/or backward scattering. Note that since the interaction region is finite, momentum need not be conserved in the scattering process. We have Hint = H0 = H′ = v ¯ (5a) (5b) (5c) √ B d (x, y) = φ1 (x) A2 (y) + A† (y) / π + 1 ↔ 2. 2 1 [∂x Φ(x)]2 g2 ¯ 1 [∂x Θ(x)]2 , g2 ¯ (9) dx aP (x)Π(x) + b ∂x Φ(x)∂x Θ(x) , g2 ¯ (10) 2 2 where v = (v1 + v2 ), 1/¯2 = (v1 /g1 + v2 /g2 )/4¯, a = ¯ g v 2 2 2 2 (v1 − v2 )/¯, and b = (v1 /g1 − v2 /g2 )/(v1 /g1 + v2 /g2 ). v The current operator for the LL model is given by √ ji = vF Pi / π and through the continuity equation, the √ current is expressed as jj (x) = −∂t φj (x, t)/ π. Using the Kubo formula, we obtain the transconductance in terms of the new fields defined in Eq. (7) as where B (x, y) = A1 (x)A2 (y) + h.c., B c (x, y) = [∂x φ1 (x)∂y φ2 (y)] /π, dx [P (x)]2 + and where H ′ describes the interaction between the new field operators B a (x, y) + B b (x, y) + B c (x, y) + B d (x, y) , (4) b v ¯ 2 + [Π(x)]2 + dxdy U12 (x, y) × B a (x, y) = A1 (x)A† (y) + h.c., 2 (8) (5d) G21 (ω) = iωe2 r r [DΦ (x, x′ ; ω) − DΘ (x, x′ ; ω)] , 4π (11) where the Green’s functions DΦ (t − t′ ) = −iΘ(t − t′ ) [Φ(x, t′ ), Φ(x′ , t)] , Here A(x) is a ”backscattering operator” defined as √ A(x) = exp (i2kF x + i2 πφ(x)) /(2πα). The two terms B b and B d correspond to nonmomentum-conserving scatterings processes. The terms B c and B d do not provide a mechanism for Coulomb drag (to any order in perturbation theory) and, furthermore, since the renormalization due to these terms is not important at low energies, they are omitted. We will make one further approximation which is valid when the relevant energy scale is smaller than vF /LI . In this limit, the spatial dependence of φ’s in the backscattering operator DΘ (t − t′ ) = −iΘ(t − t′ ) [Θ(x, t), Θ(x′ , t′ )] , (12a) (12b) have been defined (note that D(x, x′ ; ω) is independent of x, x′ in the dc-limit). For identical wires the part of the Hamiltonian which couples the Φ and Θ sector in Eq. (10) is equal to zero. The remaining Hamiltonian is equivalent to two LLs models scattering against single impurities, but with new interaction parameters. We can therefore use the results from this well-studied problem [18,19]. The transconductance simplifies to 2 G21 = 1 (GLutt (V1 , 2g) − GLutt (V2 , 2g)) , 4 For a long interaction region, i.e. W1 ≪ 1, G21 approaches the value e2 /2h for small temperatures (but still larger than T ∗ = D|W1 |−2/γ ), see inset of Fig. 2. In this regime, the diagonal conductance [21] also tends to e2 /2h and thus the currents in two wires become the same, which is similar to the absolute drag effects found in Ref. [14]. This interesting effect occurs because the momentum conserving backscattering increases with decreasing temperature (second term of Eq. (13) goes to zero). It saturates when the two currents are the same and the net momentum exchange hence is zero. (13) where g ≡ g1 = g2 and where GLutt (V, 2g) [20] is the conductance of a LL with interaction parameter, 2g, scattering against a single impurity with a backscattering amplitude V = V (2kF ). (Below it is shown that Eq. (13) is valid even when the velocities are different.) Here we have defined V1,2 = D 2πvF dxdy U12 (x, y) exp [2ikF (x ± y)] , (14) where the small momentum cut-off is parametrized in terms of a high energy cut-off: D = vF /α. Now several conclusions follow. Firstly, it is seen that for a short interaction region LI kF ≪ 1, which implies V1 ≈ V2 , all momentum transfers have equal weights and hence there is no Coulomb drag [21]. More importantly, the scaling properties of the model can be read out [18,19]: For g > 1/2, GLutt goes to a constant as T → 0, which means that G21 goes to zero at zero temperature. The corrections to the low temperature limit give the power law: G21 ∼ T 4g−2 . For non-interacting wires the drag effect is thus quadratic in temperature; in contrast to the case where they interact throughout the wires, in which case the drag scales linearly with temperature [13]. For g < 1/2, both terms in Eq. (13) go to zero, because the model scales toward strong backscattering. The power laws are however the same but the prefactors will be different and we can conclude that G21 ∼ T 1/g−2 . For g = 1/2 the problem maps to Fermi-liquids and we get a temperature independent G21 and hence only for g = 1/2 does the transconductance remain finite as temperature goes to zero. Consider again the case of identical wires in the vicinity of g = 1/2, where the problem maps to that of two Fermi-liquids. We may use the perturbative renormalization method developed in Ref. [22] rather than the exact solution based on the Bethe ansatz [19], and we obtain for the transconductance [20] G21 (|W2 |2 − |W1 |2 )tγ e2 , = 4π [1 + |W1 |2 tγ ][1 + |W2 |2 tγ ] FIG. 2. Transconductance for two identical wires using the expression derived for g close to one half. The full (dashed) curves are for T /D = 0.1 (0.001). For a short interaction region (left panel, where kF LI = 1) the transconductance peaks near g = 1/2. The peak moves toward g = 1/2 for smaller temperature. For long interaction region (right panel, where kF LI = 20), the peak is not developed and instead G21 approaches e2 /2h at low temperature for g < 1/2, which is shown in the inset for g = 0.4. This regime corresponds to the limit where only momentum conserving scattering is possible and where the currents in two wires become almost the same. In the general case, where the wires have different gvalues and velocities, there is a coupling term in the transformed Hamiltonian, Eq. (10). Since the conductance involves only the fields at x = 0, we integrate out all the x = 0 fields. The resulting action reads S = S0 + Sint , 1 |ωn | ∗ S0 = (Φω Θω ) β ω g ¯ (15) n where t = (T /D), Wi = Vi /vF , and γ = 4g − 2. In Fig. 2, we show G21 expressed in Eq. (15) for different parameters. The interwire interaction U12 is calculated for typical parameters for GaAs quantum wires and the distance between the wires is chosen to be 2/kF . In accordance with the arguments given above, the transconductance peaks when g = 1/2 at low temperatures and the peak moves to smaller g values for higher temperatures. (Furthermore, the stronger the interwire interaction the closer is the peak position to g = 1/2.) Therefore for g < 1/2, G21 shows non-monotonic behavior as a function of temperature. β Sint = √ πΘ(τ ) dτ V2 e2i (16a) k+ k− k− k+ √ + V1 e2i Φω Θω πΦ(τ ) , (16b) + c.c. , (16c) 0 1/2 1 2 2 . where k± = 2 (v2 + v1 )(g1 ± g2 )2 /(g1 v2 + g2 v1 ) For the special case, when the two wires have the same interaction parameter (but unequal velocities) the two sectors separate and the solution is again given by Eq. (13). For the general situation, we perform a perturbative renormalization group calculation and find dVi = (1 − (g1 + g2 ))Vi . d ln D 3 (17) The flow is marginal when the sum of the g-constants is equal to one and the zero temperature drag predicted above occurs at this line and for g1 + g2 < 1 a strong enhancement of the Coulomb drag occurs. Finally, we consider Coulomb drag of edge excitations in the FQH regime. Recently, there has been a large activity trying to understand the low energy edge excitations and the tunneling experiments [4] in FQH states [6,7]. These works show that the edge excitations can be described as unidirectional bosonic edge excitations, which were then used to form an electron charge creation operator as input in an independent boson model for the tunneling density of states. [1] I. Safi and H. J. Schulz, Phys. Rev. B 52, R17040 (1996) D. I. Maslov and M. Stone, Phys. Rev. B 52, R5539 (1995); V. Ponomarenko, Phys. Rev. B 52, R8666 (1995); A. Kawabata, Phys. Soc. Jap. 65, 30 (1996); Y. Oreg and A. Finkel’stein, Phys. Rev. B 54, R14265 (1996). I. Safi, Phys. Rev. B 55 R7331 (1997). [2] S. Tarucha, T. Honda, and T. Saku, Solid State Comm., 94 (1995) 413; K. Thomas et al., Phys. Rev. Lett. 77, 135 (1996); A. Yacoby et al., Phys. Rev. Lett. 77, 4612 (1996). [3] X. Wen, Int. J. Mod. Phys. B 6, 1711 (1992). [4] M. Grayson et al., Phys. Rev. Lett. 80, 1062 (1998), and references therein. [5] I. Aleiner and L. Glazman, Phys. Rev. Lett. 72, 2935 (1995). [6] S. Conti and G. Vignale, Phys. Rev. B 54, R14309 (1996), cond-mat/9709055, cond-mat/9801318; J. Han and D. Thouless, Phys. Rev. B 55, R1926 (1997); J. Han, Phys. Rev. B 56, 15806 (1997); U. Z¨licke and A. H. MacDonu ald, cond-mat/9802019. [7] D.V. Khveshchenko, cond-mat/9710137. [8] K. Flensberg and B. Y.-K. Hu, Phys. Rev. Lett. 73, 3572 (1994); N. P. R. Hill et al., Phys. Rev. Lett. 78, 2204 (1997). [9] T. J. Gramila et al., Phys. Rev. Lett. 66, 1216 (1991), Phys. Rev. B 47, 12957 (1993), Physica B 197, 442 (1994); M.C. Bønsager et al., Phys. Rev. B, in press. [10] N.P.R. Hill et al., J. Phys.: Condens. Matter 5, 5009 (1993); H. Rubel et al., Phys. Rev. Lett. 78, 1763 (1997); E. Shimshoni and S.L. Sondhi, Phys. Rev. B 49, 11484 (1994); M.C. Bønsager et al., Phys. Rev. Lett. 77, 1366 (1996), Phys. Rev. B 56, 10314 (1997). [11] M.P. Lilly, J.P. Eisenstein, L.N. Pfeiffer and K.W. West, Phys. Rev. Lett. 80, 1714 (1998). [12] S. Sakhi, Phys. Rev. B 56, 4098 (1997); I. Ussishkin and A. Stern, Phys. Rev. B 56, 4013 (1997); Y.B. Kim and A.J. Millis, cond-mat/9611125. [13] B. Y.-K. Hu and K. Flensberg, in Hot Carriers in Semicondoctors, edited by K. Hess (Plenum Press, New York, 1996), p. 943. [14] Y. V. Nazarov and D. V. Averin, cond-mat/9705158. These authors consider the first term of Eq. (6) valid for long wires. [15] Note that since our effect evolves a transport experiment the zero temperature drag considered here is different from that of a closed system in studied in A. G. Rojo and G. D. Mahan, Phys. Rev. Lett. 68, 2074 (1992). [16] J. S´lyom, Advances in Physics 28, 201 (1970). o [17] The Fermion operator should include a phase factor which ensures that the electron operators corresponding to different branches anticommute. However, they are not important for the present problem. [18] C. L. Kane and M. P. A. Fisher, Phys. Rev. Lett. 68, 1220 (1992); Phys. Rev. B 46, 15233 (1992). [19] P. Fendley, A. Ludwig, and H. Saleur, Phys. Rev. B 52, 8934 (1995). Following these works, we write the Hamiltonian that governs the dynamics for (say) the left edge branch as Hedge,L = 2π vD k>0 ρL (k)ρL (−k), where ρL is the L left branch 1D charge operator and [ρL (−k), ρL (k)] = νkL/2π. Here ν is the filling factor, vD is the drift velocity at the edge and L a normalization length. Similar expression is obtained for the right branch, similar to the Tomonaga-Luttinger model. The operator which creates a single charge e at point x, moving with velocity vD , in the left channel is Ψ† ∼ L 2π e−ikD x exp νL k ρL (k)eikx /k . Notice that this operator does not fulfill the correct anticommutation relations for fermions. However, the backscattering operator A = Ψ† ΨL does obey the correct commutation relations R and below it is utilized to describe interedge tunnelings in a model Hamiltonian for coupled FQH systems. The following experiment is proposed: the coupled QH systems are narrowed by a constriction such that edge states moving in opposite directions are coupled and backscattering can occur. This system can be modeled by writing the backscattering interaction in terms of the backscattering operator, A. After some simple algebra, we arrive at the Hamiltonian for two coupled LLs, Eqs. (1) and (6) , with g = ν [23] and conclusions similar to above therefore immediately follow from Eq. (13): In particular, we predict non-zero drag at zero temperature for half-filled Landau levels and furthermore near its maxium value the drag effect increases with decreasing temperature (or voltage). For long interaction region, the transconductance goes to a universal value νe2 /2h at low temperature for ν < 1/2. In conclusion, we have calculated Coulomb drag for LLs and found interesting behavior near g=1/2 such as zero temperature drag and non-monotonic temperature dependence. These predictions can be tested for edge states in FQH systems. The author acknowledges valuable discussions with Ben Hu and Antti-Pekka Jauho. Recently, a related paper appeared [24]. These authors discuss the conductance of crossed LLs coupled at a single point and find very similar results to ours. 4 [20] Be aware that in absence of scattering GLutt (2g) = 2e2 /h, because in this case g should be set to one[1]. This is not the case for the FQH edge states. [21] The diagonal conductance is G11 given by Eq. (13) with a plus instead of the minus and G11 of course remains finite for V1 = V2 . [22] D. Yue, L. Glazman, and K. Matveev, Phys. Rev. B 49, 1966 (1994). [23] These result will change when including interedge interactions, which lead to small logarithmic corrections. See Ref. [7] and K. Moon and S. M. Girvin, Phys. Rev. B 54, 4448 (1996); U. Z¨licke and A. H. MacDonald, ibid 54, u R8349 (1996). [24] A. Komnik and R. Egger, Phys. Rev. Lett. 80, 2881 (1998). 5