A twisted ladder: relating the Fe superconductors to the high-t c. cuprates. Related content. Recent citations

Similar documents
A Twisted Ladder: relating the Fe superconductors to the high T c cuprates. (Dated: May 7, 2009)

A Twisted Ladder: Relating the Iron Superconductors and the High-Tc Cuprates

arxiv: v1 [cond-mat.supr-con] 12 Dec 2010

Nodal and nodeless superconductivity in Iron-based superconductors

Exact results concerning the phase diagram of the Hubbard Model

Anisotropic Magnetic Structures in Iron-Based Superconductors

Workshop on Principles and Design of Strongly Correlated Electronic Systems August 2010

New perspectives in superconductors. E. Bascones Instituto de Ciencia de Materiales de Madrid (ICMM-CSIC)

arxiv: v1 [cond-mat.supr-con] 25 Mar 2014

Supporting Information

More a progress report than a talk

Workshop on Principles and Design of Strongly Correlated Electronic Systems August 2010

Mott physics: from basic concepts to iron superconductors

Back to the Iron age the physics of Fe-pnictides

Pairing Symmetry in a Two-Orbital Exchange Coupling Model of Oxypnictides

Tuning order in cuprate superconductors

ORBITAL SELECTIVITY AND HUND S PHYSICS IN IRON-BASED SC. Laura Fanfarillo

arxiv:cond-mat/ v2 [cond-mat.str-el] 27 Dec 1999

Effect of next-nearest-neighbour interaction on d x 2 y2-wave superconducting phase in 2D t-j model

Supplementary Figures.

Магнетизм в железосодержащих сверхпроводниках: взаимодействие магнитных, орбитальных и решеточных степеней свободы

ORBITAL SELECTIVITY AND HUND S PHYSICS IN IRON-BASED SC. Laura Fanfarillo

A New Electronic Orbital Order Identified in Parent Compound of Fe-Based High-Temperature Superconductors

Spin correlations in conducting and superconducting materials Collin Broholm Johns Hopkins University

Physics of iron-based high temperature superconductors. Abstract

Space group symmetry, spin-orbit coupling and the low energy effective Hamiltonian for iron based superconductors

Origin of the superconducting state in the collapsed tetragonal phase of KFe 2 As 2 : Supplemental Information

2D Bose and Non-Fermi Liquid Metals

arxiv: v1 [cond-mat.supr-con] 27 Feb 2017

Topological order in the pseudogap metal

Nematic quantum paramagnet in spin-1 square lattice models

Magnetism and Superconductivity in Decorated Lattices

edited by Nan-Lin Wang Hideo Hosono Pengcheng Dai MATERIALS, PROPERTIES, AND MECHANISMS IRON-BASED SUPERCONDUCTORS

Properties of the multiorbital Hubbard models for the iron-based superconductors

The underdoped cuprates as fractionalized Fermi liquids (FL*)

Superconductivity in Fe-based ladder compound BaFe 2 S 3

arxiv: v2 [cond-mat.supr-con] 11 Jul 2018

J. Phys.: Condens. Matter 10 (1998) L159 L165. Printed in the UK PII: S (98)90604-X

ANISOTROPIC TRANSPORT IN THE IRON PNICTIDES

Nematic and Magnetic orders in Fe-based Superconductors

From ( 0) magnetic order to superconductivity with ( ) magnetic resonance in Fe 1.02 (Te 1-x Se x )

Quasiparticle dynamics and interactions in non uniformly polarizable solids

Computational strongly correlated materials R. Torsten Clay Physics & Astronomy

arxiv: v2 [cond-mat.supr-con] 11 Dec 2016

arxiv:cond-mat/ v2 [cond-mat.supr-con] 1 Dec 2005

Structural and magnetic phase diagram of CeFeAsO 1-x F x and. its relationship to high-temperature superconductivity

Neutron scattering from quantum materials

The Gutzwiller Density Functional Theory

Understanding the Tc trends in high Tc superconductors. Kazuhiko Kuroki

Spin or Orbital-based Physics in the Fe-based Superconductors? W. Lv, W. Lee, F. Kruger, Z. Leong, J. Tranquada. Thanks to: DOE (EFRC)+BNL

Orbital-Selective Pairing and Gap Structures of Iron-Based Superconductors

DT I JAN S S"= = 11111'11 I HtI IBlIIIt g ~II. Report: ONR Grant N J Unclassified: For General Distribution LECTF

Spin or Orbital-based Physics in the Fe-based Superconductors? W. Lv, W. Lee, F. Kruger, Z. Leong, J. Tranquada. Thanks to: DOE (EFRC)+BNL

Let There Be Topological Superconductors

Stripes developed at the strong limit of nematicity in FeSe film

Quantum Spin-Metals in Weak Mott Insulators

FROM NODAL LIQUID TO NODAL INSULATOR

Simultaneous emergence of superconductivity, inter-pocket scattering and. nematic fluctuation in potassium-coated FeSe superconductor., and Y.

Ideas on non-fermi liquid metals and quantum criticality. T. Senthil (MIT).

Evolution of superconducting gap anisotropy in hole-doped 122 iron pnictides

Magnetic Order versus superconductivity in the Iron-based

Dao-Xin Yao and Chun Loong

Density matrix renormalization group study of a three- orbital Hubbard model with spin- orbit coupling in one dimension

Koenigstein School April Fe-based SC. review of normal state review of sc state standard model new materials & directions

Quantum simulations, adiabatic transformations,

ARPES studies of Fe pnictides: Nature of the antiferromagnetic-orthorhombic phase and the superconducting gap

Numerical Studies of the 2D Hubbard Model

arxiv: v2 [cond-mat.supr-con] 26 Oct 2008

Quantum disordering magnetic order in insulators, metals, and superconductors

Quantum phase transitions in Mott insulators and d-wave superconductors

arxiv: v1 [cond-mat.supr-con] 5 Jun 2011

The Hubbard model in cold atoms and in the high-tc cuprates

Gapless Spin Liquids in Two Dimensions

Quasi-1d Frustrated Antiferromagnets. Leon Balents, UCSB Masanori Kohno, NIMS, Tsukuba Oleg Starykh, U. Utah

Luttinger Liquid at the Edge of a Graphene Vacuum

a-. 2m ar, MICROSCOPIC DERIVATION OF THE GINZBURG-LANDAU EQUATIONS IN THE THEORY OF SUPERCONDUCTIVITY {- _? (_? - iea (r) \) 2 + G (x x')

Spin-wave dispersion in half-doped La3/2Sr1/2NiO4

Resonating Valence Bond point of view in Graphene

Chapter 7. Summary and Outlook

Electronic structure of correlated electron systems. Lecture 2

Finite-temperature properties of the two-dimensional Kondo lattice model

A Dirac Spin Liquid May Fill the Gap in the Kagome Antiferromagnet

Emergent Frontiers in Quantum Materials:

File name: Supplementary Information Description: Supplementary Notes, Supplementary Figures and Supplementary References

Examples of Lifshitz topological transition in interacting fermionic systems

A New look at the Pseudogap Phase in the Cuprates.

Pairing symmetry in iron based superconductors

Superconductivity from repulsion

Superconductivity due to massless boson exchange.

Quasi-1d Antiferromagnets

Visualizing the evolution from the Mott insulator to a charge-ordered insulator in lightly doped cuprates

How to model holes doped into a cuprate layer

arxiv:cond-mat/ v1 26 Jan 1994

Origin of the anomalous low temperature upturn in resistivity in the electron-doped cuprates.

Simple Explanation of Fermi Arcs in Cuprate Pseudogaps: A Motional Narrowing Phenomenon

Entanglement in Many-Body Fermion Systems

A DCA Study of the High Energy Kink Structure in the Hubbard Model Spectra

The phase diagram of the cuprates and the quantum phase transitions of metals in two dimensions

Which Spin Liquid Is It?

Magnetism and Superconductivity on Depleted Lattices

Transcription:

A twisted ladder: relating the Fe superconductors to the high-t c cuprates To cite this article: E Berg et al 2009 New J. Phys. 11 085007 View the article online for updates and enhancements. Related content - Near-degeneracy of several pairing channels in multiorbital models for the Fe pnictides S Graser, T A Maier, P J Hirschfeld et al. - Gap symmetry and structure of Fe-based superconductors P J Hirschfeld, M M Korshunov and I I Mazin - Sensitivity of the superconducting state and magnetic susceptibility to key aspects of electronic structure in ferropnictides A F Kemper, T A Maier, S Graser et al. Recent citations - Fractional charge and emergent mass hierarchy in diagonal two-leg t J cylinders Yi-Fan Jiang et al - Superconductivity at the border of electron localization and itinerancy Rong Yu et al - Superconductivity in a single-layer alkalidoped FeSe: A weakly coupled two-leg ladder system Wei Li et al This content was downloaded from IP address 148.251.232.83 on 03/07/2018 at 13:55

New Journal of Physics The open access journal for physics A twisted ladder: relating the Fe superconductors to the high-t c cuprates E Berg 1,3, S A Kivelson 1 and D J Scalapino 2 1 Department of Physics, Stanford University, Stanford, CA 94305-4045, USA 2 Department of Physics, University of California, Santa Barbara, CA 93106-9530, USA E-mail: erez.berg@gmail.com New Journal of Physics 11 (2009) 085007 (9pp) Received 7 May 2009 Published 24 August 2009 Online at http://www.njp.org/ doi:10.1088/1367-2630/11/8/085007 Abstract. We construct a two-leg ladder model of an Fe-pnictide superconductor and discuss its properties and relationship with the familiar two-leg cuprate model. Our results suggest that the underlying pairing mechanism for the Fe-pnictide superconductors is similar to that for the cuprates. An important question has been raised by the discovery of high-temperature superconductivity (HTS) in the Fe-pnictides: is there a single general mechanism of HTS that operates, albeit with material specific differences, in both the Fe pnictides and the cuprates (and possibly other novel superconductors), or do the cuprates and the Fe pnictides embody two of possibly many different mechanisms of HTS? This question is complicated by one of the perennial issues of the field: to what extent is it possible to understand the properties of the strongly correlated electron fluid in the Fe pnictides, the cuprates and other materials from a weak or strong-coupling perspective, given that the materials exhibit some features that appear more natural in one limit and some that are suggestive of the other. The Fe pnictides, like the cuprates, are likely in the intermediate coupling regime [1] [4], which is difficult to treat theoretically. In the present paper, we introduce a model of a two-leg ladder with an electronic structure chosen to reproduce particular momentum cuts through the Fe-pnictides band structure. We study the magnetic and superconducting properties of this model numerically using the density matrix renormalization group (DMRG) method [5] that allows us to treat the intermediate coupling problem essentially exactly. Although the existence of short range, spin gapped, antiferromagnetic correlations in the half-filled two-leg Hubbard ladder and the power-law d-wave-like pairing correlations in the doped ladder were initially unexpected [6], this behavior is now well understood [7] [9]. Thus, it is not surprising that a two-leg ladder model of 3 Author to whom any correspondence should be addressed. New Journal of Physics 11 (2009) 085007 1367-2630/09/085007+09$30.00 IOP Publishing Ltd and Deutsche Physikalische Gesellschaft

2 xy xz J 1 J 2 (a) yz xz xy yz (b) Figure 1. (a) Heisenberg model showing local Fe spins coupled by comparable nearest-neighbor J 1 and next nearest-neighbor J 2 exchange couplings. (b) Fiveorbital tight-binding Fermi surfaces from [16] with the main orbital contributions indicated: d xz (solid line), d yz (dashed line) and d xy (dotted line). The arrows illustrate the type of scattering processes that give rise to pairing in the fluctuation exchange calculations. the Fe-superconductors exhibits short range antiferromagnetic correlations and power-law pairing. However, it is interesting to see how such a model captures particular aspects of the physical properties of the Fe superconducting materials [10], specifically their spin and pairing correlations. We find that, in the limit of zero temperature, this model has a diverging superconducting susceptibility with a pair structure of a form which is the quasi-one-dimensional (1D) version of the structures which have been proposed on the basis of both weak [11] [18] and strong [19] coupling calculations for the 2D system. We also find that the dominant magnetic correlations are stripe-like with wave vector (0, π), reminiscent of the magnetic structure that is seen in the undoped parent compounds (and sometimes coexisting with superconductivity) in the Fe-pnictide superconductors. In addition, we show that the Fe-pnictide ladder is a twisted version of the two-leg Hubbard ladder. Upon untwisting the ladder, a correspondence is found with the familiar results from cuprate related studies: the superconductivity in the Fe ladder corresponds to d-wave-like correlations in the cuprate Hubbard ladder, and the stripe-like magnetic correlations are transformed into antiferromagnetic correlations with wave vector (π, π). This supports the idea that there is a single, unified mechanism at work in these two families of HTSs. As in the case of the cuprates, both strong-coupling and weak-coupling models have been proposed to account for the magnetic, structural and superconducting properties of the Fe pnictides. From a strong-coupling perspective, the observed magnetic and structural phase transitions and much of the dynamical magnetic structure observed even in the superconducting phase are thought of as arising from frustrated quantum magnetism [20, 21]. Here, as shown in figure 1(a), one has for the undoped system a Heisenberg model in which the second-neighbor antiferromagnetic interaction, J 2, is comparable to or larger than the nearest-neighbor exchange

3 interaction, J 1. For the doped system, one has a t J 1 J 2 model. Then, as discussed in [19], a mean-field analysis shows that the dominant pairing channel involves intra-orbital (d xz,, d xz, ) and (d yz,, d yz, ) pairs, combined to form an A 1g superposition. A similar pairing structure was found in [22] in the large U limit of an exact diagonalization study of a 8 8 cluster. While both of these studies only took into account the d xz and d yz orbitals, their finding concerning the importance of intra-orbital pairing is relevant to the model we will discuss. Alternatively, the weak-coupling picture begins by considering the energy bands and the Fermi surface. Figure 1(b) shows the Fermi surfaces for a five-orbital tight-binding fit [16] of a density function theory (DFT) calculation [23], neglecting the effects of dispersion in the third direction, perpendicular to the Fe-pnictide planes. Here the color indicates the relative weights of the 3d (d xz, d yz, d xy ) orbitals that form the main contribution to the Bloch wavefunctions on the Fermi surfaces. The d xz and d yz orbitals provide the dominant weight on the hole Fermi surfaces α 1 and α 2 around the Ɣ point, whereas on the β 1 and β 2 electron Fermi surfaces one has (d yz, d xy ) and (d xz, d xy ) contributions, respectively. From a weak-coupling perspective, it is the nesting properties of the α and β Fermi surfaces that give rise to the spin density wave (SDW) instability that accounts for the magnetism of the undoped parent compounds. For the doped system, fluctuation exchange [12, 13, 16, 24, 25] and renormalization group [14, 15] calculations find that magnetic fluctuations near (π, 0) and (0, π) lead to a pairing instability. In this case, similar to the strong-coupling model, the pairing arises from the type of scattering processes illustrated in figure 1(b). Here, a (k, k ) pair on the d xz region of the α 1 Fermi surface is scattered by a Q = (0, π) spin fluctuation to a (Q + k, Q k ) pair on the d xz part of the β 2 Fermi surface. Similar (π, 0) processes involving d yz pair scattering occur between α 1 and β 1. Thus, both strong-coupling and weak-coupling approaches have been used to describe these materials. However, what is needed is an approach that allows one to treat the intermediate coupling regime. Here, using DMRG, we address this issue for a caricature of the original problem which focuses on the d xz orbital α 1 β 2 scattering process for k y = 0 and k y = π states near the Fermi surface. These scattering processes can be described by the two-leg ladder shown in the inset of figure 2 with a Hamiltonian given by H = t 1 d i,ασ d i+1,ασ 2t 2 d i,1σ d ( ) i,2σ 2t 3 d i,1σ d i+1,2σ + d i,2σ d i+1,1σ + h.c. iασ +U i,ασ n i,α n i,α. iσ iσ (1) Here, α = 1, 2 is the leg index, σ =, is the spin index, there are leg t 1, rung t 2 and diagonal t 3 one-electron hopping matrix elements and an on-site Coulomb interaction U. The factors of 2 in front of t 2 and t 3 take into account the periodic boundary conditions transverse to the ladder. The hopping strengths were fitted to the DFT band dispersion of the α 1 and β 2 pockets calculated in [23], cut through k y = 0 and π near the Fermi surfaces. At these points in momentum space, the Bloch wavefunctions have a d xz character. The parameters we used were t 1 = 0.32 and t 3 = 0.57, measured in energy units in which t 2 = 1. The ladder bandstructure is shown in figure 2, where it is compared to the DFT dispersions near the Fermi energy. The dotted line marks the location of the chemical potential for the half-filled, one electron per d xz orbital case. To capture the intermediate coupling aspect of the physics, we will take the onsite Coulomb interaction U = 3. This value of the interaction is smaller than the overall bandwidth of the ladder, but larger than the Fermi energies of the hole and electron pockets relative to their values at k x = 0.

4 4 3 2 1 ε (k x,k y ) 0 1 2 3 4 5 k y =0,DFT k y = π, DFT k y =0,Fit k y = π, Fit 6 0 0.2 0.4 0.6 0.8 1 k x /π t 2 t 3 t 1 Figure 2. Band structure ε(k x, k y ) of the ladder with t 1 = 0.32, t 2 = 1 and t 3 = 0.57. The solid and dashed curves correspond to k y = 0 and π, respectively. The black dotted line corresponds to the Fermi energy for a filling of one electron per d xz orbital. The dots represent the bands from a DFT calculation [23] for the 2D lattice, cut through k y = 0 and π. The inset shows the ladder with the one-electron hopping matrix elements t 1, t 2 and t 3. A major approximation performed here is neglecting the multi-orbital nature of the system, and the d yz orbital in particular. Consequently, the present model cannot discriminate between the A 1g (s-wave) and B 1g (d-wave) pairing symmetries, which depends on the relative sign of the pairing amplitude of the d xz and d yz orbitals. However, weak- [16] and strong- [19] coupling studies indicate that the pairing originates mostly from intra-orbital (d xz d xz and d yz d yz ) scattering processes, such as the one shown in figure 1(b). Therefore, we believe that a simple model with only the d xz orbital still captures some of the physics of the pairing mechanism. DMRG calculations were carried out for a 32 2 system. As expected, there is a spin gap, and the doped system exhibits power-law pairing correlations. At half-filling, in the presence of an externally applied magnetic field on the first site of the lower leg, we find the striped spin pattern shown in figure 3(a). The spin correlations decay exponentially with a correlation length of about four sites and a spin gap s = 0.14. For the lightly hole-doped case, n = 0.94 (4 holes), we find a spin gap s = 0.07 and power-law pairfield correlations. Here, a pairfield boundary term ( ) H 1 = 1 P 1 + h.c., (2) with 1 = 0.5 and P 1 = (d 1,1 d 1,2 d 1,1 d 1,2 ) (3) was added to the Hamiltonian. This acts as a proximity coupling to a rung on the end of the ladder. The two-leg ladder is in a phase with a single gapless charge mode, and the long-range correlations are power laws characterized by the Luttinger parameter K c. For K c > 1 the 4

0.036 ± 0.001 5 (a) (b) 2 log[ P(l x ) ] 2.5 3 3.5 4 4.5 5 5.5 0.032 ± 0.001 6 6.5 ~0 7 0 0.5 1 1.5 2 2.5 3 3.5 4 log(l x ) Figure 3. (a) Spin structure for the undoped ladder, for U/t 2 = 3. The lengths of the arrows indicate the measured values of S z (l x ) in a calculation in which an external magnetic field has been applied to the lower leftmost site. S z (l x ) decays exponentially with a correlation length of about four sites. (b) Pairfield structure for the n = 0.94 doped ladder. Here, a pairfield boundary condition has been applied to a rung on one end of the ladder and the induced rung pairfield P lx versus l x is shown on a log log plot. The amplitude of the singlet pairfield across a rung and along a diagonal at l x = 10 are shown in the inset. The error bars in the pair fields were estimated from the extrapolation of the DMRG truncation error to zero. The pairfield amplitude along a leg is less than 10 3. superconducting susceptibility is divergent, and for K c > 1 it is dominant over the charge 2 density wave susceptibility. In addition, if K c > 1, the boundary pairing term of equation (2) 2 is relevant under a renormalization group flow [26, 27] and the effective boundary condition is described by perfect Andreev reflection. In this case, the induced pair field decays as ( L tan πlx ) (1/4K c ) 2L, where L is the length of the system 4. For the stated parameters, the induced expectation value of the rung pair amplitude P (l x ) = d l x,1 d l x,2 d l x,1 d l x,2 was measured throughout the system, and is shown in figure 3(b) on a logarithmic scale. The slope of the 4 This result can be derived using an effective Hamiltonian similar to the one used in [28], with a Dirichlet boundary condition on one boundary and a Neumann boundary condition on the other (i.e. the superconducting phase is fixed on one edge, and its derivative is zero on the other).

6 curve log( P (l x ) ) versus log(l x ) gives K c 0.5, so the ladder is near the border of the superconducting (K c > 0.5) phase. This conclusion is supported by calculations with longer systems (up to 64 2). For electron doping, we found similar results with slightly smaller values of K c. The inset of figure 3(b) shows the amplitude for removing a singlet pair from two sites at a position ten sites away from the boundary. In order to interpret these results, we perform a Bardeen Cooper Schrieffer (BCS) mean-field treatment of the ladder. In such a treatment, the amplitude for removing a singlet pair from two sites separated by (l x, l y ) is A ( l x, l y ) di+lx, j+l y d i, j d i+lx, j+l y d i, j = 1 N k (k) E (k) eik l. (4) Here E (k) = [ε (k) µ] 2 + (k) 2, N is the number of sites, and the prime on the k sum implies that it is cut off when ε(k) µ > ω 0, where ω 0 is a cutoff energy. We assume that the gap function has a simple form in which (k) is approximately constant for k in the neighborhood of a given Fermi point, but with a sign change between the k y = 0 and π bands. A reasonable estimate of the relative magnitudes of the gaps on the two bands is obtained by requiring that, due to the strong onsite Coulomb interaction, the onsite pair amplitude vanishes: A (0, 0) 2 k y =0,π N ( [ ] ) ( ) 2ω 0 0, k y ky ln ( ) = 0. (5) k y Here N(0, k y ) and (k y ) are the density of states at the Fermi energy and the gap function, respectively, of the k y = 0 and π bands. We have assumed that (k y ) ω 0 W where W is the bandwidth. From equations (4) and (5), it follows then that the ratio of the amplitude for removing a singlet pair from a diagonal bond to removing it from a rung, A(1, 1)/A(0, 1), is approximately 1 (cos k 2 F (0) + cos k F (π)) 0.9. Here k F (0) and k F (π) are the k x Fermi momenta for k y = 0 and π, respectively. Similarly, the ratio of the leg to rung amplitude A(1, 0)/A(0, 1) is approximately 1 (cos k 2 F (0) cos k F (π)) 0.02. These amplitude ratios are in overall agreement with the numerical results shown in the inset of figure 3(b). (A(1, 0) is less than 10 3, which is essentially zero within our numerical accuracy.) The magnitude of the pairing amplitude depends on the strength of the boundary proximity pair field 1. Note that the rung and diagonal pairfields are in phase with each other. A similar pairfield pattern was found in weak-coupling spin-fluctuation calculations [16]. In this case, there were also d yz d yz orbital pairfield correlations, rotated by 90. The relative phase of the d xz d xz and d yz d yz pairfields determines whether the order parameter has A 1g or B 1g symmetry. The present study cannot address this issue since it lacks the d yz orbitals 5. At first sight the striped (0, π) spin structure of the undoped ladder and the pairing amplitudes of the doped ladder appear different from the (π, π) spin and d x 2 y2-like correlations familiar for the ladders used to model the cuprates. However, a closer look at the present model reveals that it is in fact identical to that used to model the cuprates. The mapping between the Fe model we have been discussing and the usual two-leg Hubbard ladder is illustrated in figure 4. First, every other rung is twisted (figure 4(a)), 5 The pairing symmetry is determined by the 2D bandstructure as well as the intra-orbital Coulomb and exchange interactions. See [16] and also [29].

7 Figure 4. (a,b) Mapping of the FeAs ladder, equation (1) to the Hubbard ladder used to model the cuprates. (a) Twist every other rung as indicated. (b) Change the sign of the d xz orbital on the shaded sites. (c) The (0, π) short range spin correlations of the FeAs ladder (shown in figure 3(a)) become the familiar (π, π) spin correlations of the usual Hubbard ladder after the twists and orbital sign changes shown in (a) and (b). (d) The pairing correlations shown in the inset of figure 3(b) are mapped to the familiar cuprate ladder d x 2 y 2-like pairfield. interchanging the two sites at the ends of the rung and leading to a ladder having a hopping t 3 along the legs and the weaker hopping t 1 along the diagonals. Then, as illustrated in figure 4(b), the phases of the orbitals on the shaded sites are changed by 1. This leads to the usual Hubbard model with t leg = 2t 3 = 1.14, t rung = 2t 2 = 2 and a next-nearest neighbor t = t 1 = 0.32. After the twist and orbital phase changes, one finds the familiar (π, π) spin correlations shown in figure 4(c) and the d x 2 y2-like structure for the pairing correlations shown in figure 4(d). Finally, we note that the parameters for the FeAs ladder place it near a region of enhanced pairing for the corresponding Hubbard ladder case [8]. There, it was found that the pairing correlations are enhanced when the parameters are such that the Fermi level is located near the top of the bonding band and the bottom of the anti-bonding band (t rung /t leg 2). The proximity to this point also makes the pairing correlations sensitive to the bandstructure parameters. We have found that while using different bandstructure parameters and a different U does not change the behavior found here qualitatively (e.g. the existence of a spin gap and the pair structure is robust), the strength of the long-range pairing correlations is very sensitive. For instance, for the bandstructure parameters used in the present paper, the pairing correlations are reduced upon increasing the interaction to U = 4. For this case, we get K c 0.3. In addition, if we use the band parameters t 1 = 0.77, t 2 = 1, t 3 = 0.65, obtained from a fit to a two-orbital model [12], and U = 4.6, we get a significant enhancement of the pairing correlations, and K c 0.75. The sensitivity of the pairing correlations to the band parameters and to U may be an artifact of the 1D nature of the ladder 6. We believe, however, that the basic superconducting mechanism in the ladder is operative in the FeAs superconductors. 6 For instance, in the ladder there is a square-root singularity of the non-interacting density of states at the van Hove points, which is known to enhance superconductivity when the chemical potential is close to these points. This causes an increased sensitivity to the band parameters and to the doping.

8 In summary, this analysis began by constructing a ladder model for the Fe pnictides with hopping parameters chosen to fit DFT calculations. Following the results of the RPA spin-fluctuation calculations and the mean-field t J 1 J 2 and exact diagonalization studies that indicate that the dominant pairing involves (0, π) d xz d xz and (π, 0) d yz d yz pair scattering processes, only the d xz orbital was kept. Using the DMRG method, we found short range magnetic correlations that mimicked the striped (0, π) magnetic order in the undoped parent system, and pairing correlations with a structure that agreed with both the weak-coupling RPA calculations and the strong-coupling results. The Fe ladder turns out to be simply a twisted version of the usual cuprate Hubbard ladder, with parameters near the regime of enhanced pairing. We believe that this is not an accident but arises from the close connection between the basic physics responsible for high-t c superconductivity in the cuprates and the Fe pnictides. Acknowledgments We thank I Affleck and S Raghu for useful discussions. DJS acknowledges the Center for Nanophase Materials Science, which is sponsored at Oak Ridge National Laboratory by the Division of Scientific User Facilities, US Department of Energy and thanks the Stanford Institute of Theoretical Physics for their hospitality. SAK was supported by the NSF through DMR 0758356. EB was supported by the US Department of Energy under contract DE-FG02-06ER46287 through the Geballe Laboratory of Advanced Materials at Stanford University. References [1] Haule K, Shim J H and Kotliar G 2008 Phys. Rev. Lett. 100 226402 [2] Si Q and Abrahams E 2008 Phys. Rev. Lett. 101 076401 [3] Si Q, Abrahams E, Dai J and Zhu J-X 2009 arxiv:0901.4112 unpublished [4] Qazilbash M M, Hamlin J J, Baumbach R E, Zhang L, Singh D J, Maple M B and Basov D N 2009 unpublished [5] White S 1992 Phys. Rev. Lett. 69 2863 [6] Dagatto E, Riera J and Scalapino D 1992 Phys. Rev. B 45 5744 [7] Dagatto E and Rice T 1996 Science 271 618 [8] Noack R, Bulut N, Scalapino D and Zacher M 1997 Phys. Rev. B 56 7162 [9] Balents L and Fisher M 1996 Phys. Rev. B 53 12133 [10] Mazin I and Schmalian J 2009 arxiv:0901.4790 unpublished [11] Mazin I, Singh D, Johannes M and Du M 2008 Phys. Rev. Lett. 101 057003 [12] Raghu S, Qi X-L, Liu C-X, Scalapino D and Zhang S-C 2008 Phys. Rev. B 77 220503 [13] Kuroki K et al 2008 Phys. Rev. Lett. 101 087004 [14] Wang F, Zhai H, Ran Y, Vishwanath A and Lee D 2009 Phys. Rev. Lett. 102 047005 [15] Chubukov A, Efremov D and Eremin I 2008 Phys. Rev. B 78 134512 [16] Graser S, Maier T, Hirschfeld P and Scalapino D 2009 New J. Phys. 11 025016 [17] Cvetkovic V and Tešanović Z 2009 Europhys. Lett. 85 37002 Cvetkovic V and Tešanović Z 2008 arxiv:0808.3742 [18] Yu S-L, Kang J and Li J-X 2009 Phys. Rev. B 79 064517 Yao Z-J, Li J-X and Wang Z D 2009 New J. Phys. 11 025009 [19] Seo K, Bernevig B A and Hu J-P 2008 Phys. Rev. Lett. 101 206404

9 [20] Fang C, Yao H, Tsai W-F, Hu J-P and Kivelson S A 2008 Phys. Rev. B 77 224509 [21] Xu C, Qi Y and Sachdev S 2008 Phys. Rev. B 78 134507 [22] Moreo A, Daghofer M, Riera J A and Dagotto E 2009 Phys. Rev. B 79 134502 [23] Cao C, Hirschfeld P J and Cheng H-P 2008 Phys. Rev. B 77 220506 [24] Ikeda H 2008 J. Phys. Soc. Japan 77 2008 [25] Yanagi Y, Yamakawa Y and Ono Y 2008 J. Phys. Soc. Japan 77 123701 [26] Affleck I, Caux J-S and Zagoskin A 2000 Phys. Rev. B 62 1433 [27] Feiguin A, White S, Scalapino D and Affleck I 2006 arxiv:cond-mat/0612636 unpublished [28] White S R, Affleck I and Scalapino D J 2002 Phys. Rev. B 65 165122 [29] Kuroki K, Usui H, Onari S, Arita R and Aoki H 2009 arxiv:0904.2612