Electron-electron interactions in a one-dimensional quantum wire spin filter P. Devillard,1 A. Cr´pieux,2 K. I. Imura,3 and T. Martin2 e arXiv:cond-mat/0501145v3 [cond-mat.mes-hall] 6 Jun 2005 2 1 Centre de Physique Th´orique, Universit´ de Provence, Case 907, 13331 Marseille Cedex 03, France e e Centre de Physique Th´orique, Universit´ de la M´diterran´e, Case 907, 13288 Marseille Cedex 9, France e e e e 3 Condensed Matter Theory Laboratory, RIKEN, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan The combined presence of a Rashba and a Zeeman effect in a ballistic one-dimensional conductor generates a spin pseudogap and the possibility to propagate a beam with well defined spin orientation. Without interactions transmission through a barrier gives a relatively well polarized beam. Using renormalization group arguments, we examine how electron-electron interactions may affect the transmission coefficient and the polarization of the outgoing beam. PACS numbers: 71.70.Ej, 72.25.Dc, 72.25.Mk Over the last decade, spintronics1 has emerged from mesoscopic physics and nanoelectronics as a field with implications in both quantum information theory2 and for the storage of information. While the charge is routinely manipulated in nanoelectronics, the issue here is to exploit the spin degree of freedom of electrons. In particular, spin filters are therefore needed to control the input and the output of spintronic devices. A recent proposal3 explained the operation of a spin filter for a one dimensional wire under the combined operation of Rashba spin orbit coupling and Zeeman splitting. Yet, electronic interactions in one dimensional wires are known to lead to Luttinger liquid behavior and to the renormalization of scattering coefficients. It therefore important to inquire about the role of electronic interactions in the above mentioned spin filter. This is aim of the present paper. A decade ago, a transistor based on the controlled precession of the electron spin due to spin orbit coupling was proposed4 . Indeed, the Rashba effect5 in a semiconductor, can be modulated by a gate voltage, which controls the asymmetry of the potential well which confines the electrons. Since this proposal, many spintronic devices based on the Rashba effect have been proposed, based on a single electron picture6,7,8,9 . At the same time, onedimensional wires are now available experimentally. This has motivated several efforts10 to study the interplay between the Rashba effect and electron interactions in infinite one-dimensional wires. However, little has been said about Coulomb interactions in spintronic devices. Our starting point stems from the fact that Rashba devices are gated devices, in which the interaction are assumed to be screened and weak. We will therefore use a perturbative renormalization group treatment of the Coulomb interaction in order to address its consequence on the spin dependent transmission. This approach will be justified by estimating the strength of electron-electron interaction in single channel semiconductor devices. Here, we consider a narrow ballistic wire which is submitted to spin orbit coupling with the Rashba term being dominant. In order to obtain a spin-polarized beam of electrons propagating in one direction, a small Zeeman field is introduced3 . By confining in the y direction with lateral gates, the problem becomes unidimensional and the Hamiltonian for a one-dimensional wire reads3 : H0 = p2 α Ez ǫZ σ − B.σ, px σy + ∗ 0 2m 2B (1) where m∗ is the effective mass, Ez is the electric Rashba field perpendicular to the layer and α depends on the material used. B is the magnetic field, ǫZ is the Zeeman energy. σ0 is identity matrix and σ = (σx , σy , σz ) the usual Pauli matrices. The eigenvectors are thus products of plane waves times a spinor. The eigenstates are: 2 E = 2m∗ 2 2 2 kx ± 2 κ4 + kα kx , Z ∗ (2) ∗ where κ2 = m2 (ǫZ /2) and kα = m2 α Ez . The Rashba Z 2 2 energy is defined through Eα = 2m∗ kα . The orientation of eigenvectors is such that for E ≫ Eα , the spinor associated to the mode with larger wave-vector is directed along |↑ y . In the interval [Eα − ǫZ /2 , Eα + ǫZ /2], there is only one propagating mode (with two chiralities). The other mode is evanescent. The dispersion relation for the propagating mode is indicated in Fig. 1. The existence of a pseudogap in a given energy interval has been used in Ref. 3, to propose a spin-filtering device, where a potential step of height V1 corresponding to a gate voltage permits to shoot in the middle of the pseudogap (Fig. 1). Electron-electron interactions are taken into account, with Vint (x), the Coulomb interaction potential. Following Ref. 11 which discusses the effect of weak electron interactions in a single mode 1D wire, we use a Hartree-Fock approach followed by a poor man Anderson renormalization in energy space. The Dyson equation in Hartree-Fock approximation reads ψk (x) = φk (x) + + dy dy Gr (x, y) VH (y)ψk (y) k dz Gr (x, y)Vex (y, z)ψk (z), (3) k whith ψk (φk ) the one-electron wave function in the presence (absence) of interactions, and Gr (x, y) is the rek tarded Green function. VH (x) = dy Vint (x − y)n(y), is 2 the Hartree potential, with: 2 n | ψq (y) |2 , n(y) = n=1 |q| 0 for repulsive interactions. Any S matrix can be written as exp(−iu.λ) with u an 8-dimensional real vector and the components of λ are ˜ ˜ ˜1,1 ˜2,1 real matrices (SU (3)). Since t1,1 , t1,2 , t′ and t′ are generally complex numbers, our renormalization equations give 8 relations for real parameters and thus completely determine the flow of the S matrix. We now focus on the limit η → 0 (small Zeeman coupling). The S matrix then simplifies and becomes a real and symmetric matrix: √ √ Γ−1   (Γ−1)2 2 Γ −2 Γ Γ+1 Γ+1     √ S = (Γ + 1)−1  2 Γ 0 Γ − 1  . (11)     √ Γ−1 Γ −2 Γ Γ+1 Γ − 1 4 Γ+1 0 0 2 4 6 8 10 A Ln T in /T (14) FIG. 3: Total transmission and polarization as a function of the logarithm of the temperature. The two curves marked with left arrows represent the total transmission for Γ = 1.01 (solid line) and Γ = 10 (dashed line). The two other curves represent the spin polarization for Γ = 1.01 (dash-dotted line) and Γ = 10. (dashed line) respectively. There are only 3 fixed points. The stable fixed point x = 0, y = 0 corresponds to a perfectly reflecting step; both t1,1 and t1,2 are zero. The unstable fixed point x = y = 1 corresponds to t1,1 = 1 and t1,2 = 0, i.e. perfect transmission with no mode conversion. The second unstable fixed point x = 1, y = −1 corresponds to t1,2 = −1 and t1,1 = 0 (complete mode conversion). When we have no electron-electron interactions, the location of the points where one starts the renormalization procedure, in the plane (t2 , t2 ), depends on one pa1,2 1,1 rameter only, Γ. Thus, the ensemble of points where one 1 (1 ideally). We then start from an initial temperature (bandwidth) Tin = D0 /kB , where t1,1 is close to 1 and t1,2 close to zero, which means a perfect polarization of the outgoing beam. The point moves first in a direction such that the polarization diminishes significantly but the total transmission remains approximately constant. Then, the curve bends downwards and eventually moves towards the origin (no transmission). The polarization goes to a constant value. In the second case (not desired in practice), Γ is large, the initial polarization is already low (equal to 2/Γ) and √ √ dx = −A(x2 − y 2 ) 1 − x 1 − 1 − x , dl dy 1 = − A(x2 − y 2 )y . dl 2 (13) 4 still decreases towards a constant value. The transmission goes quickly to zero. The renormalization procedure must be interrupted at some stage given by D ≃ kB T . In the case where Γ is close to 1 (small potential step), Eqs. (13) and (14) show that the total transmission first decreases very slowly as 2 2 1 − rin − 8rin A ln2 T /Tin , where rin is the initial reflexion coefficient. rin goes to zero as Γ goes to 1. The polarization decreases as ln(Tin /T ). For lower temperatures, the polarization goes to a constant Γ dependent value, which is equal to 1/2 for Γ = 1. The total trans−1 mission goes to zero, asymptotically like ln(Tin /T ) 2 , but this regime is only attained for unrealistic values of temperatures. A characterization of such a spin filter requires the comparison of both the total transmission coefficient, together with the evolution of the polarization under renormalization group flow. This information is illustrated in Fig. 3 where both quantities are plotted as a function of the logarithm of the inverse temperature. For Γ close to 1, the total transmission stays constant and close to unity until the temperature is dropped by several orders of magnitude, and then decreases in a monotonous fashion. At the same time, the polarization drops faster than the total transmission, signifying that the quality of the spin filtering effect is polluted by electronic interactions before the total transmission is truly affected. For large Γ, the total transmission first decreases monotonously (linear behavior in ln T ) starting from the initial bandwidth, 1 then giving place to a slower decrease ((ln T )− 2 ). The polarization stays approximately constant in this case, but at a deceptively low value of 0.2. A useful way to quantify electron interactions is to compute the Luttinger parameter g which is expected for a gated heterostructure. To be specific we consider the geometry of Ref. 14, where the Coulomb interaction in the 1D channel is screened by the few transverse modes in the wire and by the proximity of the 2DEG and gates: V (r) ≃ (e2 /4πǫ0 ǫ)e−r/λs /r (λs ∼ 100nm, a fraction of the width of the wire, W ∼ 20nm). Averaging over the 1 2 3 4 5 6 7 Proceedings of the MS+S2002 conference, Atsugi (2002), in Towards the controllable quantum states H. Takayanagi and J. Nitta (World Scientific, 2003). Proceedings of the MS+S2004 conference, Atsugi (2004), In the light of quantum computation H. Takayanagi and J. Nitta (World Scientific, 2005). ˇ P. Stˇeda and P. Seba, Phys. Rev. Lett. 90, 256601 (2003). r S. Datta and B. Das, Appl. Phys. Lett. 56 (7), 665 (1990). E. I. Rashba, Fiz. Tverd. Tela (Leningrad) 2, 1224 (1960). [ Sov. Phys. Solid State 2, 1109 (1960)]; Y. A. Bychkov and E. I. Rashba, J. Phys. C 17, 6039 (1984). J. C. Egues, C. Gould. G. Richter and L. W. Molenkamp, Phys. Rev. B 64, 195319 (2001). V. F. Motsnyi, J. De Boeck, J. Das, W. Van Roy, G. Borghs, E. Goovaerts, and V. I. Safarov, Appl. Phys. Lett. lateral dimensions of the wire, one obtains an effective one-dimensional potential. The Luttinger parameter g is then related to the zero-momentum Fourier transform ˜ of this potential g = (1 + 4V (0)/ vF )−1/2 . g increases with the ratio W/λs . Taking vF ∼ 106 m/s one obtains g ≃ 0.69, which is remarkably close to the value of Ref. 14 It is also reasonably close to the non-interacting value 1. The materials used in Ref. 14 differ from the existing Rashba devices15 , but the typical parameters are comparable. To summarize, we have looked at the effect of weak electron electron interactions in one dimensional ballistic quantum wires under the combined Rashba and Zeeman effects. At the single electron level this device has the advantage of working as a spin filter. We have characterized the influence of electron electron interactions in this same device. We found that in the most relevant case (EF −Eα ≪ 1) where the total transmission remains close to unity for a range of bandwidths, the quality of spin filtering properties decreases substantially. Although the present work deals with a sharp step, the present approach can be extended to steps whose extension is much larger than the Fermi wave length, using WKB-type approximations. Transmission through the step is likely to be enhanced in this case, but the tendency for interactions to spoil filtering will remain valid. Also, this single electron picture can be further complicated in practice: for the case where two spin orientations are present (on the right hand side of the step), it was shown16 that the combined effect of spin orbit coupling and strong electron electron interactions provide some limitations to the Luttinger liquid (metallic) picture, giving rise to spin or charge density wave behavior instead. Nevertheless, the present perturbative treatment of the Coulomb interaction is justified here because the one dimensional wire is surrounded by nearby metallic gates in order to implement the Rashba effect. We thank P. Stˇeda and F. Hekking for useful discusr sions and comments. This work is supported by a CNRS A.C. Nanosciences grant. 8 9 10 11 12 13 14 15 81, 265 (2002). E. I. Rashba, Phys. Rev. B 62, R16267 (2000). J. Schlieman and D. Loss, Phys. Rev. B 68, 165311 (2003). A. V. Moroz, K. V. Samokhin and C. H. W. Barnes, Phys. Rev. Lett. 84, 4146 (2000); Phys. Rev. B 62, 16900 (2000); A. Iucci, Phys. Rev. B 68, 075107 (2003). K. A. Matveev, Dongxiao Yue, and L. I. Glazman, Phys. Rev. Lett. 71, 3351 (1993); Dongxiao Yue, L. I. Glazman, and K. A. Matveev, Phys. Rev. B 49, 1966 (1994). M. B¨ttiker, Y. Imry, and M. Ya Azbel, Phys. Rev. A 30, u 1982 (1984). C. L. Kane and M. P. A. Fisher, Phys. Rev. Lett. 72, 724 (1994). O.M. Auslaender et al., Science 295, 825 (2002). J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki, Phys. 5 16 Rev. Lett. 78, 1335 (1997). V. Gritsev, G. Japaridze, M. Pletyukhov, and D. Baeriswyl, cond-mat/0409678.