Transport and scattering in inhomogeneous quantum wires N. Sedlmayr,1, ∗ J. Ohst,1 I. Affleck,2 J. Sirker,1, 3 and S. Eggert1, 3 arXiv:1204.2565v2 [cond-mat.str-el] 7 Sep 2012 2 1 Department of Physics, University of Kaiserslautern, D-67663 Kaiserslautern, Germany Department of Physics and Astronomy, The University of British Columbia, Vancouver, BC V6T 1Z1, Canada 3 Research Center OPTIMAS, University of Kaiserslautern, D-67663 Kaiserslautern, Germany (Dated: July 22, 2018) We consider scattering and transport in interacting quantum wires that are connected to leads. Such a setup can be represented by a minimal model of interacting fermions with sudden changes in interaction strength and/or velocity. The inhomogeneities generally cause relevant backscattering, so it is a priori unclear if perfect ballistic transport is possible in the low temperature limit. We demonstrate that a conducting fixed point surprisingly exists even for large abrupt changes, which in the considered model corresponds to a velocity matching condition. The general position dependent Green’s function is calculated in the presence of a sudden change, and is confirmed numerically with high accuracy. The exact form of the interference pattern in the form of density oscillations around inhomogeneities can be used to estimate the effective strength of local backscattering sources, offering a route to design experiments where the effects of the contacts are minimized. PACS numbers: 73.63.Nm, 71.10.Pm, 73.40.-c The description of transport through quantum wires has become a very well studied research area, as it directly ties together conductivity experiments1–4 with central theoretical models in one dimension, such as the Landauer formalism5 and Luttinger liquids.6 Landauer showed that the conductance of a single spinless noninteracting quantum channel is always finite and given by G = (1 − R2 )e2 /h, where R is the backscattering amplitude.5 However, interaction effects change this picture dramatically as R becomes temperature dependent through renormalization. In the pioneering work of Kane and Fisher it was found that a local perturbation in a repulsive Luttinger liquid is relevant, which results in a characteristic powerlaw dependence R ∝ T g−1 where g < 1 is the interaction parameter.7 A number of impurity models in one dimension have since been analyzed in detail8–16 and confirm that generically a local perturbation cuts the transport at low temperatures, unless the relevant operator is forbidden by symmetry. Nearly perfect connections between leads and wires can be achieved for different types of quantum wires,1–4 which show quantized conductance at moderate temperatures. At lower temperatures, however, scattering in the connections to the leads plays an increasingly important role. The higher dimensional contacts are effectively weakly interacting, so there has been great interest in describing the transport through a quantum wire attached to noninteracting leads.16–28 The minimal model for this setup is a single channel of interacting spinless fermions, where the interaction parameter changes along the wire. Under the assumption that it is possible to use a hydrodynamic Luttinger liquid description for this model,17 it would be expected that backscattering follows a non-trivial renormalization behavior which is position dependent.18 Even if the connections are effectively free from imperfections with a homogeneous lattice structure,4 there have to be small regions of the wire, the junctions, where the interaction changes, which does not necessarily occur adi- abatically and will induce intrinsic backscattering. This immediately invites the question if it is ever possible to create a perfect connection in the low temperature limit. Indeed it is in general unclear how large such intrinsic backscattering is for generic junctions and if it is even justified to use an inhomogeneous Luttinger liquid description in the first place for this setup, since the usual bosonization procedure assumes a translational and scale invariant theory. In this paper we consider the intrinsic backscattering in an inhomogeneous wire and show that a perfectly conducting fixed point can indeed be found by adjusting parameters such as the velocities on both sides. At half filling it is also possible to estimate the bare strength of the backscattering from the change in the local velocity field. The full correlation function of an inhomogeneous Luttinger liquid is calculated, which agrees with numerical simulations to high accuracy. This confirms the validity of the inhomogeneous theory and proves that a low energy conducting fixed point can exist even in the presence of large abrupt jumps. A hydrodynamic description of interacting fermions with a changing interaction parameter gx leads to a generalized inhomogeneous Luttinger liquid action17,18 1/T dτ S0 = 0 dx ux (∂τ φ)2 + (∂x φ)2 , 2gx u2 x (1) where φ is a canonical bosonic field. Here the effective velocity ux generally also changes with the interaction strength. We will derive the corresponding correlation function for an abrupt jump below. Additionally, however, backscattering must be considered. In particular, it is known that the Hamiltonian density contains a leading relevant oscillating operator given by7 √ 4πφ † e−i2kF x ψ+ ψ− ∝ e−i2kF x e−i , (2) where ψ± are left and right moving fermion fields and kF is the Fermi-wavenumber. Normally this operator 2 can be neglected under the integral, but this is no longer the case for inhomogeneous systems. In fact, it is exactly this operator as a local perturbation which causes the renormalization of defects in wires7 and spin chains.8 The oscillating part of the interaction results in the same bosonic operator,8 so that even a change in interaction alone will induce backscattering. In principle conductance depends on the connections of the wire to both of the leads. We will concentrate here on the backscattering in the case where the length of the wire is larger than the coherence length ∝ u/T so that it suffices to consider each junction separately.18 A single junction can be characterized by changes taking place in a small region around x = 0. The field φx is assumed to be slowly varying on the scale of the Fermi wavelength so that it is possible to use an expansion of φx in Eq. (2) and a partial integration to derive an effectively local perturbation √ H ′ ≈ λe−i 4πφx=0 + h.c., Dφe−S0 − 1/T dτ H ′ 0 (4) with the quadratic action in Eq. (1). For an abrupt change in the interaction parameter at x = 0 from gx<0 = gℓ to gx≥0 = gr and similarly for the velocity ux the gen˜ eral bosonic Green’s function G(x, x′ ; τ ) = φx,τ φx′ ,0 0 is found by matching the left and right parts. In particular, it is possible to solve ˜ G(x, x′ ; τ ) = T ˜ eiωm τ Gm (x, x′ ) (5) m with 2 ωm ux ∂ ∂ − 2gx ux ∂x 2gx ∂x ˜ Gm (x, x′ ) = δ(x − x′ ) g ¯ ˜ G(x, x′ ; τ ) = − ln sinh πT π (6) |x| |x′ | + − iτ ux ux′ sinh πT L[x, x′ ]gx ln + π sinh πT |x| ux + |x′ | ux′ |x−x′ | ux (7) − iτ , − iτ where we have defined a new effective interaction param1 eter g = 2( gℓ + g1 )−1 . Here L[x, x′ ] is 1 when x and x′ ¯ r are in the same region, and 0 when they are not. The renormalization of a local perturbation in Eq. (3) can be determined with the help of the Green’s function by integrating out the Fourier components above a cutoff Λ,6,18 which gives 1 dλ = 1 − g. ¯ λ d ln Λ (3) while the oscillating operator cancels everywhere in the uniform regions. So far we have kept only the leading relevant operator, but there are higher order terms which will be discussed later. In general λ is complex, except for particle hole symmetric situations. Even relatively smooth junctions become effectively sharper and sharper under renormalization, so that the relevant backscattering is non-zero unless the amplitude λ is adjusted to zero, which requires the fine tuning of two parameters. As we will see later it is indeed possible to identify such conducting fixed points in a lattice model by an appropriate choice of parameters on both sides of the junction. The existence of a conducting fixed point is also of relevance for the discussion about possible charge fractionalization in Luttinger liquids.3,29,30 If the Luttinger liquid supports only charges e(1 ± gx )/2 related to the chiral eigenstates of the Hamiltonian on each side of the wire, then it seems to be impossible that backscattering can be tuned to zero. Our results thus support the analysis in Ref. [30] that at such a junction an arbitrary charge can be injected into a Luttinger liquid. In order to perform a renormalization group analysis we have to consider the full partition function Z= ˜ by allowing a discontinuity in the derivative of Gm at ′ 17 x = x . We determine the general Green’s function to be (8) We therefore expect that the effective backscattering renormalizes as a power law in the temperature R ∝ ¯ T g−1 . As a concrete lattice model we can consider spinless fermions at half-filling H= x † 1 1 −tx (ψx ψx+1 + h.c.) + Ux (nx − 2 )(nx+1 − 2 ) , (9) † where nx = ψx ψx . The corresponding interaction parameter g and the renormalized velocity u in Eq. (1) are functions of U and t, which are known from the Bethe ansatz.6 For small jumps and interactions we have estimated the size of λ perturbatively, which turns out to be proportional to the corresponding renormalized velocity field ux λ∝ x e−i2kF x (ux+1 − ux ), (10) with 2kF = π, so to lowest order it does not matter if the velocity change occurs due to inhomogeneous interactions or hopping amplitudes, which may even compensate each other. In particular, Eq. (10) suggests that a conducting fixed point can be achieved for a sharp discontinuity by a velocity matching uℓ = ur , but the derivation above is valid only for small interactions. The numerical simulations show that this condition holds even in a strongly interacting model. Equation (10) also implies that backscattering can be made arbitrarily small by very slow ”adiabatic” changes. In order to calculate the backscattering amplitude R numerically, we use quantum Monte Carlo simulations31 on systems with an abrupt junction at x = 0, i.e. Ux<0 = Uℓ and Ux≥0 = Ur and analogously for the hopping amplitude tx . For long system sizes L 40tx /T the boundary condition at ±L/2 becomes irrelevant. The backscattering R induces a 2kF interference pattern in the density, the so-called Friedel oscillations,13,16,32 which can 3 0.2 0.12 0.45 χ g-1 ( T/TK) left side right side R(T) 0.4 g-1 0.11 (T/TK) left side right side R(T) 0.1 0.09 0.35 0.1 0.08 0.3 0 -40 0.07 0.06 0.25 -30 -20 -10 0 10 20 30 x 40 (a) PSfrag replacements Uℓ = 0 non-interacting Ur = 1.8tr interacting FIG. 1: (Color online) The local response in the density χ for a jump in interaction from Uℓ = 0 to Ur = 1.8tr at T = 0.1tr . Circles (blue): No discontinuity in hopping, tℓ = tr . Squares (red): The hopping on the left is adjusted to tℓ ≈ 1.518tr in order to match the velocities on both sides. Solid lines (black) are fits to the predicted behavior in Eq. (14) for even and odd sites separately. be calculated directly in the simulations and also give additional information about the correlation functions. In particular, we consider the local oscillating density in a half-filled lattice in response to changing the chemical potential χx = ∂ nx ∂µ , (b) 0.01 (11) µ=0 0.1 T/t analogously to the local magnetization and susceptibility in spin chains.8,13,32–35 The results for the Friedel oscillations χ = χ0 + (−1)x χalt near an abrupt junction at x = 0 are shown in Fig. 1 for a change from Uℓ = 0 to Ur = 1.8tr at finite temperatures T = 0.1tr for two cases: In the case that the interaction strength changes but the hopping is equal tℓ = tr , we have uℓ < ur and strong alternating 2kF = π oscillations are observed (circles). In the second case, the hopping tℓ on the left was also increased in order to exactly match the velocities 2tℓ = uℓ = ur ≈ 3.036tr (squares). Clearly, the backscattering oscillation is strongly suppressed with a different position dependence, but it is not zero. As the numerical data will show it turns out that the relevant backscattering term is exactly zero in this case. The uniform parts of χx on the two sides are constant and approximately given gx by the zero temperature result χ0 = πux , which are not equal in either case. We have also tested other cases with more complicated changes in hopping and interactions over three sites in order to tune λ = 0 in Eq. (10), and in all cases backscattering is strongly suppressed. Let us first analyze the position dependence of the Friedel amplitude in the case of unequal velocities, which 0.1 shows a characteristic maximum in Fig. 1 reminiscent of the local susceptibility near ends in spin chains.32,35 However, we will show that the behavior is not exactly of the same form as for scattering from open ends as was conjectured before.13 In order to calculate the alternating response in the presence of the perturbation (3) with small |λ| ∝ |ur − uℓ | we use the bosonized form of the density (12) 1/T √ (−1) 1 sin[ 4πφx ] (12) nx = n0 − √ ∂x φx + const. π π T/t FIG. 2: (Color online) Effective backscattering amplitude R(T ) on a logarithmic scale extracted from the amplitude of the density oscillations in Eq. (14) for a jump from (a) Uℓ = 0 to Ur = 1.8t, (TK = 3.4 × 10−4 t), and (b) Uℓ = 1t to Ur = 1.4t with t = tℓ = tr , (TK = 6 × 10−6 t). The amplitudes are extracted from fitting the local response for x < 0 (“left side”) and x > 0 (“right side”) separately. where the density is bosonized as x 0.05 0.01 χalt = λ dτ 0 √ √ ∂ sin 4πφx,τ cos 4πφ0,0 ∂µ 0 , (13) where the dependence on µ is given by a shift of the µgx field ∂x φ by √πux which turns the sine-dependence into a cosine-dependence with an additional factor of x. Using the Green’s function in Eq. (7) we arrive 1 at integrals of the form 0 dτ | sinh(X − iπτ )|−2g = 2g (sinh 2X)−g P−g (coth 2X) where Pl (z) is the Legendre function. Therefore we find χalt ∝ λ gx x ux √ √ dτ cos 4πφx,τ cos 4πφ0,0 1/T 0 ¯ ∝ λT g −1 x ux 2πT x sinh T ux −gx P−¯ (z) g 0 (14) where z = coth[2πT x/ux ]. Compared to the scattering from an open boundary13,32 there are two notable differences: first of all there is an additional factor in the form of the Legendre function, which changes the shape significantly near the scatterer, but quickly approaches unity for x ux /πT . Second, there is an additional temper¯ ature dependent factor ∝ T g −1 which is in agreement with the renormalization behavior predicted in Eq. (8). As can be seen by the fit in Fig. 1, the behavior (14) describes the numerical data perfectly, where x is taken to be the position from an effective scattering center. We 4 χalt uℓ = 0.98ur uℓ = ur uℓ = 1.02ur 0.01 0 PSfrag replacements 0 10 20 30 x 40 FIG. 3: (Color online) The alternating part of the density oscillations at T = 0.1tr on the interacting side U = 1.8tr , where the hopping on the non-interacting side is adjusted so that uℓ = 0.98ur , uℓ = ur , and uℓ = 1.02ur , respectively. have confirmed the position dependence for many different temperatures and discontinuities. This unambiguously shows that the action in Eq. (1) and the Green’s function (7) leading up to Eq. (14) are a reliable description of the problem. An interesting detail in Fig. 1 is the maximum in the non-interacting region where Uℓ = 0 and gℓ = 1 which is absent for an open end but now arises due to the Legendre function with g < 1. Physically this can be understood ¯ as a proximity effect, where the behavior in a range on the non-interacting side is influenced by the collective excitations on the other side. The temperature dependent amplitudes of the fits to the position-dependent part in Eq. (14) directly give the backscattering R(T ) in Fig. 2, where we have used the amplitude from open ends corresponding to R = 1 as the normalization, which is known exactly.35,36 The data confirm the predicted power-law behavior at low temperatures. Note that the independent fits on both sides give roughly the same R(T ). For the larger jump in the interaction U = 1.8tr there are deviations at higher temperatures coming from higher order operators (left panel). Below a characteristic temperature TK we expect the power law to break down as the stable fixed point R(0) = 1 is approached, but this energy scale could not be reached in the simulations. In order to find a conducting scenario in the low temperature limit, it is interesting to analyze the case of equal velocities uℓ = ur with strong discontinuities in both hopping and interaction in more detail. As shown in Fig. 1 density oscillations are still observed in this case, but we are interested only in the contribution of the leading relevant operator in Eq. (3), which will grow while the temperature is lowered and must change sign as a function of velocity at the conducting fixed point. In Fig. 3 we show the alternating part of the oscillations for a jump from Uℓ = 0 to Ur = 1.8tr , where the hopping on the left has been adjusted to three different cases: uℓ = 0.98ur , uℓ = ur , and uℓ = 1.02ur . It is quite apparent that a sign change takes place exactly at uℓ = ur , which means that the relevant backscattering vanishes. The remaining oscillations visible in Fig. 1 for uℓ = ur are caused by higher order local operators in Eq. (3).√ In particular, the next-leading terms are given √ ¯ by ∂x ei 4πφ and ei 16πφ with scaling dimension g + 1 and 4¯, respectively, which are irrelevant unless 4¯ < 1. g g Away from half filling the marginal operator ∂x φx=0 will also be present,12 but does not affect the scattering to first order. The velocity matching rule for a conducting fixed point is surprisingly simple, considering that it is not linked to any special symmetry in this scenario. Perfect conduction in quantum wires with impurities has been described before in cases where a renormalization to the periodic boundary condition fixed point occurs.7 This, however, is not the case here. We find a novel conducting fixed point described by Eq. (7) which does not correspond to a standard boundary conformally invariant theory. To conclude, we have studied the intrinsic backscattering present when connecting a quantum wire to a lead. All inhomogeneities and impurities in the junction are in general relevant for repulsive Luttinger liquids, leading to a conductance which scales as G(T ) ∝ 2¯−2 g e2 ], with an unusual power-law exponent h [1 − (T /TK ) 1 given by g = 2( gℓ + g1 )−1 in terms of the interaction pa¯ r rameters on both sides of the junction gℓ , gr . We have shown that it is possible to achieve a perfect connection, i.e. an absence of relevant backscattering, in inhomogeneous wires by tuning other parameters such as the velocity in the lead and in the wire. The general Green’s function was calculated in the presence of an abrupt jump along the wire based on an inhomogeneous Luttinger liquid action, which was in excellent agreement with numerical simulations. The results suggest that the observation of Friedel oscillations in the density along the wire, which can be attempted by scanning probe microscopy, can be used to analyze local scattering centers. A systematic study of inhomogeneities and thus an experimental test of our results seems feasible in semi-conductor heterostructures. One could start from a device where a quantum wire adiabatically broadens into a two-dimensional electron gas.4 By applying a sharp potential barrier at the interface a local backscattering center could then be realized allowing it to continuously tune the backscattering amplitude λ. Acknowledgments We are thankful for discussions with J. Folk. This work was supported by the DFG via the SFB/Transregio 49, NSERC, CIfAR, and the MAINZ (MATCOR) graduate school of excellence. 5 ∗ 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 Electronic address: sedlmayr@physik.uni-kl.de W. Liang, M. Bockrath, D. Bozovic, J.H. Hafner, M. Tinkham, and H. Park, Nature 411, 665 (2001); A. Javey, J. Guo, Q. Wang, M. Lundstrom, and H. Dai, Nature 424, 654 (2003). A. Yacoby, H.L. Stormer, N.S. Wingreen, L.N. Pfeiffer, K.W. Baldwin, and K.W. West, Phys. Rev. Lett. 77, 4612 (1996) H. Steinberg, G. Barak, A. Yacoby, L.N. Pfeiffer, K.W. West, B.I. Halperin and K. Le Hur, Nature Phys. 4, 116 (2008). S. Tarucha, T. Honda, and T. Saku, Sol. State Comm. 94, 413 (1995). R. Landauer, IBM J. Res. Dev. 1 (1957), 223. For a review, see T. Giamarchi, Quantum physics in onedimension, Oxford, 2003. C.L. Kane and M.P.A. Fisher, Phys. Rev. B 46, 15233 (1992); Phys. Rev. Lett. 68, 1220 (1992); Phys. Rev. B 46, 7268 (1992). S. Eggert and I. Affleck, Phys. Rev. B 46, 10866 (1992). A. Furusaki and N. Nagaosa, Phys. Rev. B 47, 4631 (1993). E.S. Sørensen, S. Eggert, and I. Affleck, J. Phys. A 26, 6757 (1993). P. Fr¨jdh and H. Johannesson, Phys. Rev. B 53, 3211 o (1996). S.J. Qin, M. Fabrizio, L. Yu, M. Oshikawa, and I. Affleck, Phys. Rev. B 56, 9766 (1997). S. Rommer and S. Eggert, Phys. Rev. B 62, 4370 (2000). S. Eggert and S. Rommer, Phys. Rev. Lett. 81, 1690 (1998); S. Eggert, D.P. Gustafsson, and S. Rommer Phys. Rev. Lett. 86, 516 (2001). R.G. Pereira and E. Miranda, Phys. Rev. B 69, 140402 (2004); N. Sedlmayr, S. Eggert, and J. Sirker, Phys. Rev. B 84, 024424 (2011). D. Yue, L.I. Glazman, and K.A. Matveev, Phys. Rev. B 49, 1966 (1994). 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 D.L. Maslov and M. Stone, Phys. Rev. B 52, R5539 (1995). A. Furusaki and N. Nagaosa, Phys. Rev. B 54, R5239 (1996). M. Ogata and H. Fukuyama, Phys. Rev. Lett. 73, 468 (1994). E. Wong and I. Affleck, Nucl. Phys. B 417, 403 (1994). V.V. Ponomarenko, Phys. Rev. B 52, R8666 (1995). I. Safi and H.J. Schulz, Phys. Rev. B 52, R17040 (1995). C.C. Chamon and E. Fradkin, Phys. Rev. B 56, 2012 (1997). K.-I. Imura, K.-V. Pham, P. Lederer, and F. Pi´chon, e Phys. Rev. B 66, 035313 (2002). T. Enss, V. Meden, S. Andergassen, X. BarnabeTheriault, W. Metzner, and K. Sch¨nhammer, Phys. Rev. o B 71, 155401 (2005); K. Janzen, V. Meden, and K. Sch¨nhammer, Phys. Rev. B 74, 085301 (2006). o J. Rech and K.A. Matveev, Phys. Rev. Lett. 100, 066407 (2008); J. Phys.: Condens. Matter 20, 164211 (2008). D.B. Gutman, Y. Gefen, and A.D. Mirlin, Europhys. Lett. 90, 37003 (2010). R. Thomale and A. Seidel, Phys. Rev. B 83, 115330 (2011). K.-V. Pham, M. Gabay and P. Lederer, Phys. Rev. B 61, 16397 (2000). K. Le Hur, B.I. Halperin and A. Yacoby, Ann. Phys. 323, 3037 (2008). O.F. Sylju˚ asen and A.W. Sandvik, Phys. Rev. E 66, 046701 (2002). S. Eggert and I. Affleck, Phys. Rev. Lett. 75, 934 (1995). J. Sirker, N. Laflorencie, S. Fujimoto, S. Eggert and I. Affleck, Phys. Rev. Lett. 98, 137205 (2007). J. Sirker, N. Laflorencie, S. Fujimoto, S. Eggert and I. Affleck, J. Stat. Mech. P02015 (2008). J. Sirker and N. Laflorencie, Europhys. Lett. 86, 57004 (2009). S. Lukyanov and V. Terras, Nucl. Phys. B 654, 323 (2003).