cond-mat0508373/all.tex
1: %\documentclass[aps,prb,showpacs,floatfix,preprint]{revtex4}
2: \documentclass[aps,prb,showpacs,twocolumn,floatfix]{revtex4}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \usepackage{graphicx}
5: \usepackage{amsmath}
6: \usepackage{amsfonts}
7: \usepackage{amssymb}
8: \usepackage{bm}
9: \usepackage{dcolumn}
10: 
11: \begin{document}
12: 
13: \title{First-principles scattering matrices for spin-transport}
14: \author{K. Xia}
15: \altaffiliation[Present address:]{State Key Laboratory for Surface
16: Physics, Institute of Physics, Chinese Academy of Sciences, P. O.
17: Box 603, Beijing 100080, China.}
18: \author{M. Zwierzycki}
19: \altaffiliation[Present address:]
20: {Max-Planck-Institut f\"{u}r Festk\"{o}rperforschung, Heisenbergstr. 1,
21: D-70569 Stuttgart, Germany.}
22: \altaffiliation[Permanent address:]{Institute of
23: Molecular Physics, P.A.N., Smoluchowskiego 17, 60-179 Pozna\'n,
24: Poland.}
25: \author{M. Talanana}
26: \author{P. J. Kelly}
27: \affiliation{Faculty of Science and Technology and MESA$^{+}$
28: Research Institute, University of Twente, P.O. Box 217, 7500 AE
29: Enschede, The Netherlands}
30: \author{G. E. W. Bauer}
31: \affiliation{Kavli Institute of NanoScience, Delft University of
32: Technology, Lorentzweg 1, 2628 CJ Delft, The Netherlands }
33: 
34: \date{\today }
35: 
36: \begin{abstract}
37: Details are presented of an efficient formalism for calculating
38: transmission and reflection matrices from first principles in layered
39: materials. Within the framework of spin density functional theory and
40: using tight-binding muffin-tin orbitals, scattering matrices are
41: determined by matching the wave-functions at the boundaries between
42: leads which support well-defined scattering states and the scattering
43: region. The calculation scales linearly with the number of principal
44: layers $N$ in the scattering region and as the cube of the number of atoms
45: $H$ in the lateral supercell. For metallic systems for which the required
46: Brillouin zone sampling decreases as $H$ increases, the final scaling goes
47: as $H^{2}N$. In practice, the efficient basis set allows scattering
48: regions for which $H^{2}N \sim 10^{6}$ to be handled.
49: The method is illustrated for Co/Cu multilayers and single interfaces
50: using large lateral supercells (up to $20\times 20$) to model interface
51: disorder. Because the scattering states are explicitly found,
52: ``channel decomposition'' of the interface scattering for clean and
53: disordered interfaces can be performed.
54: \end{abstract}
55: 
56: 
57: \pacs{72.10.Bg,72.25.Ba,75.47.De}
58:         
59: \maketitle
60: 
61: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% INTRODUCTION
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% INTRODUCTION
63: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% INTRODUCTION
64: 
65: \section{INTRODUCTION}
66: \label{sec:intro}
67: 
68: One of the most important driving forces in condensed matter physics
69: in the last thirty years has been the controlled growth of layered
70: structures so thin that interface effects dominate bulk properties and
71: quantum size effects can be observed. In doped semiconductors, the
72: large Fermi wavelength of mobile charge carriers made it possible to
73: observe finite size effects for layer thicknesses on a micron scale.
74: Much thinner layers must be used in order to make such observations in
75: metals because Fermi wavelengths are typically of the order of an
76: interatomic spacing. Nevertheless, following rapidly on the heels of a
77: number of important discoveries in semiconductor heterostructures,
78: interface-dominated effects such as interface magnetic anisotropy,
79: oscillatory exchange coupling and giant magnetoresistance (GMR) were
80: found in artificially layered transition metal materials. Reflecting
81: the shorter Fermi wavelength, the characteristic length scale is of
82: order of nanometers.
83: 
84: 
85: Our main purpose in this paper is to give details of a scheme we have
86: developed which is suitable for studying mesoscopic transport in
87: inhomogeneous, mainly layered, transition metal magnetic materials.
88: In the context of a large number of schemes designed to study transport
89: either from first-principles\cite{
90: Schep:prl95,
91: Zahn:prl95,
92: Weinberger:jp96,
93: Schep:prb97,
94: Zahn:prl98,
95: vanHoof:prb99,
96: MacLaren:prb99,
97: Kudrnovsky:prb00,
98: Xia:prb01,
99: Riedel:prb01,
100: Taylor:prb01,
101: Brandbyge:prb02,
102: Drchal:prb02,
103: Wortmann:prb02a,Wortmann:prb02b,
104: Weinberger:prp03,
105: Thygesen:prb03,
106: Mavropoulos:prb04}
107: or based upon electronic structures calculated from first-principles\cite{
108: Mathon:prb97a,
109: Mathon:prb97b,
110: Tsymbal:jp97,
111: Sanvito:prb99,
112: Velev:prb03,Velev:prb04}
113: we will require our computational scheme to be (i)
114: physically transparent, % Landauer-B\"uttiker
115: (ii) first-principles, requiring no free parameters, % DFT/LDA
116: (iii) capable of handling complex electronic structures characteristic of
117: transition metal elements and (iv) very efficient in order to be able to
118: handle lateral supercells to study layered systems with different lattice
119: parameters and to model disorder very flexibly. % TB-MTO
120: A tight-binding (TB) muffin-tin-orbital (MTO) implementation of the
121: Landauer-B\"{u}ttiker formulation of transport theory within the
122: local-spin-density approximation (LSDA) of density-functional-theory (DFT)
123: will satisfy these requirements.
124: 
125: Because wave transport through interfaces is naturally described
126: in terms of transmission and reflection, the Landauer-B\"{u}ttiker
127: (LB) transmission matrix formulation of electron transport gained
128: rapid acceptance as a powerful tool in the field of mesoscopic
129: physics,\cite{Imry:02,Datta:95} once the controversies surrounding
130: the circumstances under which different expressions should be used
131: had been resolved. \cite{Imry:02} The two-terminal conductance of
132: a piece of material is measured by attaching leads on either side,
133: passing a current through these leads and measuring the potential
134: drop across the scattering region. In the LB formulation of
135: transport theory, the conductance $G$ is expressed in terms of a
136: transmission matrix $t\equiv t(E_{F})$
137: \begin{equation}
138: G=\frac{e^{2}}{h}Tr\{tt^{\dag }\}  \label{eq:LB}
139: \end{equation}%
140: where the element $t_{\mu \nu }$ is the probability amplitude that a state
141: $\left\vert \nu \right\rangle $ in the left-hand lead incident on the
142: scattering region from the left (see Fig.~\ref{fig:LSR}) is scattered into a
143: state $\left\vert \mu \right\rangle $ in the right-hand lead.
144: The trace simply sums over all incident and transmitted ``channels''
145: $\nu $ and $\mu $ and $\frac{e^{2}}{h}$ is the fundamental unit of
146: conductance. In much current work on first-principles transport the
147: conductance is calculated directly from Green's functions expressed in
148: some convenient localized orbital representation.\cite{Fisher:prb81}
149: Explicit calculation of the scattering states is avoided by making use
150: of the invariance properties of a trace. Because we want to make contact
151: with a large body of theoretical literature\cite{Beenakker:rmp97}
152: on mesoscopic physics and address a wider range of problems in the
153: field of spin-dependent transport, we will calculate the
154: microscopic transmission and reflection matrices $t$ and $r$. By
155: using a real energy, we will avoid the problems encountered in
156: distinguishing propagating and evanescent states when a small but
157: finite imaginary part of the energy is used. The
158: Landauer-B\"{u}ttiker formalism satisfies our first requirement of
159: physical transparency.
160: 
161: \begin{figure}[tbp]
162: \includegraphics[scale=0.40,angle=0,clip=true]{LSR2a.eps}
163: \caption{Sketch of the configuration used in the Landauer-B\"{u}ttiker
164: transport formulation to calculate the two terminal conductance.
165: A (shaded) scattering region ($\mathcal{S}$) is sandwiched by
166: left- ($\mathcal{L}$) and right-hand ($\mathcal{R}$) leads which
167: have translational symmetry and are partitioned into principal layers
168: perpendicular to the transport direction. The scattering region
169: contains $N$ principal layers but the structure and chemical
170: composition are in principle arbitrary.}
171: \label{fig:LSR}
172: \end{figure}
173: In developing a scheme for studying transport in transition metal
174: multilayers, a fundamental difference between semiconductors and
175: transition metals must be recognized. Transition metal atoms have
176: two types of electrons with different orbital character. The
177: \textit{s} electrons are spatially quite extended and, in solids,
178: form broad bands with low effective masses; they conduct easily.
179: The \textit{d} electrons are much more localized in space, form
180: narrow bands with large effective masses and are responsible for
181: the magnetism of transition metal elements. The ``magnetic''
182: electrons, however, being itinerant do contribute to electrical
183: transport. The appropriate framework for describing metallic
184: magnetism, even for the late 3$d$ transition metal elements, is
185: band theory.\cite{Kubler:00} An extremely successful framework
186: exists for treating itinerant electron systems from
187: first-principles and this is the Local Density Approximation (LDA)
188: of Density Functional Theory (DFT). For band magnetism, the
189: appropriate extension to spin-polarized systems, the local
190: spin-density approximation (LSDA) satisfies our second requirement
191: of requiring no free parameters.\cite{fn:total_en}
192: 
193: Oscillatory exchange coupling in layered magnetic structures was discussed
194: by Bruno in terms of generalized reflection and transmission matrices \cite%
195: {Bruno:jmmm93} which were calculated by Stiles
196: \cite{Stiles:jap96,Stiles:prb96,Stiles:prb00} for realistic electronic structures
197: using a scheme \cite{Stiles:prb88,Stiles:prl91} based on
198: linearized augmented plane waves (LAPWs). At an interface between a
199: non-magnetic and a magnetic metal, the different electronic structures
200: of the majority and minority spin electrons in the magnetic material
201: give rise to strongly spin-dependent reflection.
202: \cite{Schep:prl95,Schep:prb98} Schep used transmission and reflection
203: matrices calculated from first-principles with an embedding surface
204: Green's function method \cite{vanHoof:97} to calculate spin-dependent
205: interface resistances for specular Co/Cu interfaces embedded in
206: diffusive bulk material.\cite{Schep:prb97} The resulting good
207: agreement with experiment indicated that interface disorder is less
208: important than the spin-dependent reflection and transmission from a
209: perfect interface.  Calculations of domain wall resistances as a
210: function of the domain wall thickness illustrated the usefulness of
211: calculating the full scattering matrix.
212: \cite{vanHoof:jmmm98,vanHoof:prb99} However, the LAPW basis set used
213: by Stiles and Schep was computationally too expensive to allow
214: repeated lateral supercells to be used to model interfaces between
215: materials with very different, incommensurate lattice parameters or to
216: model disorder. This is true of all plane-wave based basis sets which
217: typically require of order 100 plane waves per atom in order to
218: describe transition metal atom electronic structures reasonably well.
219: 
220: Muffin-tin orbitals (MTO) form a flexible, minimal basis set leading to
221: highly efficient computational schemes for solving the Kohn-Sham equations
222: of DFT.\cite{Andersen:prl84,Andersen:85,Andersen:prb86,Andersen:87}
223: For the close packed structures adopted by the magnetic materials Fe, Co, Ni
224: and their alloys, a basis set of 9 functions ({\em s, p}, and {\em d} orbitals)
225: per atom in combination with the atomic
226: sphere approximation (ASA) for the potential leads to errors in describing
227: the electronic structure which are comparable to the absolute errors
228: incurred by using the local density approximation. This should be compared
229: to typically 100 basis functions per atom required by the more accurate LAPW
230: method. MTOs thus satisfy our third and fourth requirements of being able to
231: treat complex electronic structures efficiently.
232: 
233: The tight-binding linearized muffin tin orbital (TB-LMTO) surface Green's
234: function (SGF) method has been developed to study the electronic structure
235: of interfaces and other layered systems. When combined with the
236: coherent-potential approximation (CPA), it allows the electronic structure,
237: charge and spin densities of layered materials with substitutional disorder
238: to be calculated self-consistently very efficiently.\cite{Turek:97} In this
239: paper we describe how we have combined a method for calculating transmission
240: and reflection matrices based on wave-function matching (WFM), in a form
241: given by Ando\cite{Ando:prb91} for an empirical tight-binding Hamiltonian,
242: with a first-principles TB-MTO basis. \cite{Andersen:prb86} Applications of
243: the method to a number of problems of current interest in spin-transport
244: have already been given in a number of short publications: to the
245: calculation of spin-dependent interface resistances where interface disorder
246: was modelled by means of large lateral supercells; \cite{Xia:prb01} to the
247: first principles calculation of the so-called mixing conductance parameter
248: entering theories of current-induced magnetization reversal\cite{Xia:prb02}
249: and Gilbert-damping enhancement via spin-pumping;\cite{Zwierzycki:prb05}
250: to a generalized scattering formulation of the suppression of Andreev
251: scattering at a ferromagnetic/superconducting interface; \cite{Xia:prl02} to
252: the problem of how spin-dependent interface resistances influence spin
253: injection from a metallic ferromagnet into a III--V semiconductor.\cite%
254: {Zwierzycki:prb03,Bauer:prl04} These examples amply demonstrate that the
255: fourth requirement is well satisfied.
256: 
257: In Sec.~\ref{sec:theory}, we give technical details of the formalism and
258: illustrate it in Sec.~\ref{sec:Calc} where we calculate the transmission
259: matrices for clean and disordered Co/Cu interfaces, document a number of
260: convergence and accuracy issues and give a detailed ``channel-decomposition''
261: analysis of the scattering in the presence of disorder. A comparison with
262: other methods is made in Sec.~\ref{sec:discussion}.
263: 
264: \section{THEORY}
265: \label{sec:theory}
266: 
267: Central to the wave-function matching method for calculating the
268: transmission and reflection matrices is the equation of motion (EoM) for
269: electrons with energy $\varepsilon $, relating the vectors of coefficients $%
270: \mathbf{C}_{I}$ for layers $I-1$, $I$, and $I+1$:
271: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
272: \begin{equation}
273: {\mathcal{H}}_{I,I-1}\mathbf{C}_{I-1}+({\mathcal{H}}-\varepsilon )_{I,I}%
274: \mathbf{C}_{I}+{\mathcal{H}}_{I,I+1}\mathbf{C}_{I+1}=0.  \label{eq:EoM}
275: \end{equation}%
276: Here, $\mathbf{C}_{I} \equiv C_{Ii}$ describes the wavefunction
277: amplitude in terms of some localized orbital basis $\left\vert
278: i\right\rangle $ of dimension $M$ where $i$ labels the atomic
279: orbital and atom site. [For the muffin-tin orbitals to be outlined
280: in Sec.~\ref{ssec:MTOs}, $i$ will be a combined index
281: $\mathbf{R}lm$, where $l$ and\ $m$ are the azimuthal and magnetic
282: quantum numbers, respectively, of the MTO defined for an
283: atomic-spheres-approximation (ASA) potential on the site
284: $\mathbf{R}$.] The EoM does not restrict us to only considering
285: nearest neighbour interactions since atoms can always be grouped
286: into layers defined as to be so thick that the interactions
287: between layers $I$ and $I\pm 2$ are negligible (see
288: Fig.~\ref{fig:LSR}). Such layers are called \textit{principal
289: layers}. Their thickness depends on the range of the interactions
290: which in turn partly depends on the spatial extent of the orbital
291: basis. It will be minimized by using the highly localized
292: tight-binding MTO representation.
293: 
294: Consider the situation sketched in Fig.~\ref{fig:LSR} where the
295: scattering region $\mathcal{S}$ is contacted with left ($\mathcal{L)}$
296: and right ($\mathcal{R)}$ leads which have perfect lattice periodicity
297: and support well-defined scattering states. We assume that the ground
298: state charge and spin densities and the corresponding Kohn-Sham
299: independent electron potential have already been calculated
300: self-consistently. The calculation of the scattering matrix can now be
301: split into two distinct parts. In the first stage, to be discussed in
302: Sec.~\ref{ssec:Leads}, the eigenmodes of the leads
303: $\mathbf{u}_{\mu } (= \mathbf{C}_{0}$ for the $\mu$-th mode),
304: of which there are
305: $2M$, are calculated using an EoM appropriate to MTOs and making use
306: of the lattice periodicity. By calculating their \textbf{k }vectors
307: (which are in general complex)\ and velocities $\upsilon_{\mathbf{k}}$,
308:  the eigenstates can be classified as being either left-going
309: $\mathbf{u}_{\mu }(-)$ or right-going $\mathbf{u}_{\mu }(+)$. They form a
310: basis in which to expand any left- and right-going waves and have the
311: convenient property that their transformation under a lattice translation in
312: the leads is easily calculated using Bloch's theorem (with \textbf{k }%
313: complex). We use the small Roman letters \emph{i,j} to label the
314: non-orthogonal basis and the small Greek letters $\mu ,\nu $ to label the
315: lead eigenmodes.
316: 
317: In the second stage discussed in Sec.~\ref{ssec:scatt}, a scattering
318: region $\mathcal{S}$ is introduced in the layers $1 \leq I \leq {\rm N}$
319: which mixes left- and right-going lead eigenmodes. The scattering region
320: can be a single interface, a complex multilayer or a tunnelling junction,
321: and the scattering can be introduced by disorder or simply by
322: discontinuities in the electronic structure at interfaces.
323: The $\nu \rightarrow \mu $ element of the reflection matrix,
324: $r_{\mu \nu },$ is defined in terms of the ratio of the amplitudes of
325: left-going and right-going solutions in the left lead (in layer 0 for example)
326: projected onto the $\nu ^{\text{th}}$ right-going and $\mu ^{\text{th}}$
327: left-going propagating states (\textbf{k }vector real) renormalized with the
328: velocities so as to have unit flux. The scattering problem is solved by
329: direct numerical inversion of a matrix with the leads included as a boundary
330: condition so as to make finite the matrix which has to be inverted.
331: 
332: \subsection{Muffin Tin Orbitals and the KKR equation}
333: \label{ssec:MTOs}
334: 
335: Muffin-tin orbitals
336: \cite{Andersen:prl84,Andersen:85,Andersen:prb86,Andersen:87} (MTO) are
337: defined for potentials in which space is
338: divided into non-overlapping atom-centred ``muffin-tin'' spheres inside
339: which the potential is spherically symmetric and the remaining
340: ``interstitial'' region where the potential is taken to be
341: constant. The atomic spheres approximation (ASA) is obtained (i) by taking
342: the kinetic energy in the interstitial region to be zero and (ii) by
343: expanding the muffin-tin spheres so that they fill all space whereby the
344: volume of the interstitial region vanishes; for monoatomic solids such
345: spheres are called atomic Wigner-Seitz (WS) spheres. Inside a WS (or MT)
346: sphere at $\mathbf{R}$, the solution of the radial Schr\"{o}dinger equation
347: regular at $\mathbf{R}$, $\phi _{Rl}(\varepsilon ,r_{R})$ can be determined
348: numerically for energy $\varepsilon $ and angular momentumum $l$ resulting
349: in the partial wave
350: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
351: \begin{equation}
352: \phi _{Rlm}(\varepsilon ,\mathbf{r}_{R})\equiv \phi _{RL}(\varepsilon ,%
353: \mathbf{r}_{R})\equiv \phi _{Rl}(\varepsilon ,r_{R})Y_{lm}(\mathbf{\hat{r}}%
354: _{R})  \label{eq:PWdef}
355: \end{equation}%
356: where $\mathbf{r}_{R}\equiv \mathbf{r-R}$ and \ $r_{R}\equiv \left\vert
357: \mathbf{r-R}\right\vert $. A continuous and differentiable orbital is
358: constructed by attaching \ to the partial wave at the sphere boundary
359: $r_{R}\equiv s_{R}$ a ``tail'' consisting of an appropriate linear
360: combination of the solutions of the Laplace equation,
361: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
362: \begin{equation}
363: J_{RL}^{0}(\mathbf{r}_{R})\equiv \;(r_{R}/\omega )^{l}[2(2l+1)]^{-1}Y_{L}(%
364: \mathbf{\hat{r}}_{R})  \label{eq:J0}
365: \end{equation}%
366: and
367: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
368: \begin{equation}
369: K_{RL}^{0}(\mathbf{r}_{R})\equiv (r_{R}/\omega )^{-l-1}Y_{L}(\mathbf{\hat{r}}%
370: _{R}),  \label{eq:K0}
371: \end{equation}%
372: which are respectively, regular at $\mathbf{R}$ and at infinity. $\omega $
373: is the average WS radius if the structure contains different atoms. In terms
374: of the logarithmic derivative of $\phi _{l}(\varepsilon ,r)$ at $r\equiv s$
375: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
376: \begin{equation}
377: D_{l}(\varepsilon ,s)\equiv \frac{s\phi _{l}^{\prime }(\varepsilon ,s)}{\phi
378: _{l}(\varepsilon ,s)}  \label{eq:LD}
379: \end{equation}%
380: ($\phi _{l}^{\prime }(\varepsilon ,s)$ is the radial derivative), the radial
381: solutions are matched if for $r \geqslant s$,
382: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
383: \begin{align}
384: \phi _{l}(\varepsilon ,r)& =\frac{l-D_{l}}{2l+1}\left( \frac{s}{\omega }%
385: \right) ^{l+1}\phi _{l}(\varepsilon ,s)  \notag \\
386: & \times \left[ K_{l}^{0}(r)-2(2l+1)\left( \frac{\omega }{s}\right) ^{2l+1}(%
387: \frac{D_{l}+l+1}{D_{l}-l})J_{l}^{0}(r)\right]
388: \end{align}%
389: where we drop the explicit $\mathbf{R}$-dependence when this does not give
390: rise to ambiguity, or in terms of the potential function
391: \begin{equation}
392: P_{l}^{0}(\varepsilon )=2(2l+1)
393: \left( \frac{\omega}{s}\right) ^{2l+1}
394: \frac{D_{l}(\varepsilon )+l+1}{D_{l}(\varepsilon )-l}
395: \end{equation}%
396: and normalization $N_{l}^{0}(\varepsilon )=
397: \frac{2l+1}{l-D_{l}}
398: \left( \frac{\omega }{s}\right) ^{l+1} \frac{1}{\phi _{l}(\varepsilon ,s)}$
399: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
400: \begin{equation}
401: N_{l}^{0}(\varepsilon )\phi _{l}(\varepsilon
402: ,r)=K_{l}^{0}(r)-P_{l}^{0}(\varepsilon )J_{l}^{0}(r)
403: \end{equation}
404: By subtracting from the partial wave, both inside and outside the MT sphere,
405: the $J_{RL}^{0}(\mathbf{r}_{R})$ component which is irregular at infinity, a
406: function is formed which is continuous, differentiable and regular in all
407: space, an energy-dependent muffin tin orbital $\chi _{RL}^{0}(\varepsilon ,%
408: \mathbf{r}_{R})$:
409: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
410: \begin{align}
411: \chi _{RL}^{0}(\varepsilon ,\mathbf{r}_{R})& =N_{Rl}^{0}(\varepsilon )\phi
412: _{Rl}(\varepsilon ,\mathbf{r})+P_{Rl}^{0}(\varepsilon )J_{RL}^{0}(\mathbf{r}%
413: _{R}) & r_{R}& \leqslant s_{R} \\
414: & =K_{L}^{0}(\mathbf{r}_{R}) & r_{R}& \eqslantgtr s_{R}  \label{eq:MTOdef}
415: \end{align}%
416: The tail $K_{RL}^{0}(\mathbf{r}_{R}\mathbf{)}$\ has the desirable property
417: that closed forms exist for expanding it around a different site $\mathbf{R}%
418: ^{\prime }$ in terms of the regular solutions $J_{R^{\prime }L^{\prime
419: }}^{0}(\mathbf{r}_{R^{\prime }}\mathbf{),}$%
420: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
421: \begin{equation}
422: K_{RL}^{0}(\mathbf{r}_{R}\mathbf{)=-}\sum_{L^{\prime }}
423: J_{R^{\prime }L^{\prime }}^{0}(\mathbf{r}_{R^{\prime }})
424: S_{R^{\prime }L^{\prime },RL}^{0}
425: \label{eq:TailExp}
426: \end{equation}%
427: The expansion coefficients $S_{R^{\prime }L^{\prime },RL\text{ }}^{0}$form a
428: so-called canonical structure constant matrix: they do not depend on the
429: lattice constant, on the MT (or AS) potentials or on energy. Because of the
430: augmentation with $J_{RL}^{0}(\mathbf{r}_{R})$,\ the resulting MTO is no
431: longer a solution of the Schr\"{o}dinger equation (SE) inside its own sphere
432: $\mathbf{R}$. When, however, a solution of the SE is sought in the form of a
433: linear combination of MTOs centred on different sites,
434: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
435: \begin{equation}
436: \Psi (\varepsilon ,\mathbf{r})=\sum_{R,L}\chi _{RL}^{0}(\varepsilon ,\mathbf{%
437: r}_{R})C_{RL}^{0}  \label{eq:LCMTO}
438: \end{equation}%
439: then the partial wave solution is recovered if the augmenting term $%
440: J_{RL}^{0}(\mathbf{r}_{R})$ on site $\mathbf{R}$ is cancelled by the tails
441: of MTOs centred on all other sites $\mathbf{R}^{\prime }\neq \mathbf{R}$,
442: expanded about $\mathbf{R.}$ The condition for this to occur is the
443: ``tail-cancellation'' condition:
444: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
445: \begin{equation}
446: \sum_{R^{\prime },L^{\prime }}\left[ P_{RL}^{0}(\varepsilon )\delta
447: _{RR^{\prime }}\delta _{LL^{\prime }}-S_{RL,R^{\prime }L^{\prime }}^{0}\right]
448: C_{R^{\prime }L^{\prime }}^{0}=0.  \label{eq:PmS}
449: \end{equation}%
450: All of the information about the structural geometry of the system
451: under investigation is contained in the structure constant matrix
452: $S_{RL,R^{\prime }L^{\prime }}^{0}$ while all of the information about
453: the atomic species on site $\mathbf{R}$ needed to calculate the
454: electronic structure (eigenvalues and eigenvectors) is contained in the
455: potential functions $ P_{RL}^{0}(\varepsilon )$. These are determined
456: by solving the radial Schr\"{o}dinger equation for the corresponding
457: spherically symmetrical atomic sphere potential for energy $\varepsilon$
458: and angular momentum $l$.
459: 
460: A disadvantage of these ``conventional'' MTOs is their infinite range.
461: However, there is a remarkably simple generalization of the MTOs which allows
462: their range to be modified by introducing a set of ``screening'' constants
463: $\alpha _{Rl}$ (not to be confused with the lead eigenmode index) while the
464: ``tail-cancellation'' condition remains essentially unchanged:
465: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
466: \begin{equation}
467: \sum_{R^{\prime },L^{\prime }}\left[ P_{RL}^{\alpha }(\varepsilon )
468: \delta_{RR^{\prime }}\delta _{LL^{\prime }}-S_{RL,R^{\prime }L^{\prime }}^{\alpha }%
469: \right] C_{R^{\prime }L^{\prime }}^{\alpha }=0.  \label{eq:TBPmS}
470: \end{equation}
471: $P^{\alpha }(\varepsilon )$ is a diagonal matrix related to
472: $P^{0}(\varepsilon )$ by
473: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
474: \begin{equation}
475: P^{\alpha }(\varepsilon )=P^{0}(\varepsilon )+P^{0}(\varepsilon )\alpha
476: P^{\alpha }(\varepsilon )=\frac{P^{0}(\varepsilon )}{1-\alpha
477: P^{0}(\varepsilon )},
478: \end{equation}%
479: and%
480: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
481: \begin{equation}
482: S^{\alpha }=S^{0}+S^{0}\alpha S^{\alpha }=S^{0}\left(1-\alpha S^{0}\right)^{-1},
483: \end{equation}%
484: For any set of $\alpha _{Rl}$, the energy-dependent MTOs with the
485: normalization
486: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
487: \begin{equation}
488: \sum_{R,L}\frac{\omega }{2}\dot{P}_{RL}^{\alpha }(\varepsilon )\left\vert
489: C_{RL}^{\alpha }\right\vert ^{2}=1,  \label{eq:Norm}
490: \end{equation}%
491: form a complete set for the MT (AS) potential used in their construction.
492: Here, $\dot{P}$ denotes an energy derivative and (\ref{eq:Norm}) follows
493: from the relation
494: $N^{\alpha}(\varepsilon)
495:   =  [(\omega/2) \dot{P}^{\alpha}(\varepsilon)]^{1/2}$.
496: Sets of parameters $\alpha_{Rl}$ have been found for which the ``screened''
497: structure constants $S_{RL,R^{\prime }L^{\prime}}^{\alpha }$ have very short
498: range, decaying exponentially with the interatomic separation.
499: \cite{Andersen:85} The set of parameters, $\beta _{Rl}$, which yields
500: the shortest range MTOs is called the ``tight-binding'' (TB) representation.
501: \cite{Andersen:prl84} For close-packed structures, the range of
502: $S_{RL,R^{\prime }L^{\prime }}^{\beta }$ is in practice limited to first-
503: and second-nearest neighbours. This TB set with $\alpha =\beta $ is what
504: we will use from now, unless stated otherwise, since it will allow us to
505: define principal layers with a minimal thickness.
506: 
507: For the determination of energy bands $\varepsilon (\mathbf{k}),$ the
508: tail-cancellation or KKR equations are inconvenient because the
509: energy-dependence of the potential function makes it necessary to solve (\ref%
510: {eq:PmS}) or (\ref{eq:TBPmS}) by searching for the roots of a determinant,
511: which is very time consuming. Much more efficient methods have been
512: developed based on energy-independent MTOs. However, to study transport we
513: only need to know $P^{\beta }(\varepsilon )$ for a fixed energy, usually the
514: Fermi energy. We assume that the Kohn-Sham equations have already been
515: solved self-consistently (using for example a linearized method) so we have
516: the potentials from which to calculate the potential functions. Although
517: (\ref{eq:TBPmS}) can be brought into Hamiltonian form by linearizing the
518: energy dependent potential function (see Appendix~\ref{sec:velocities}),
519: we will work directly with the more exact KKR equation.
520: 
521: 
522: \subsection{Eigenmodes of the leads}
523: \label{ssec:Leads}
524: 
525: We will assume that there exists two-dimensional translational symmetry in
526: the plane perpendicular to the transport direction so that states can be
527: characterized by a lateral wave vector ${\bf k_{\parallel }}$ in the
528: corresponding two-dimensional Brillouin zone. The screened KKR equation
529: \cite{Andersen:85} in the mixed representation of ${\bf k_{\parallel }}$
530: and real space layer index $I$ (see Fig.~\ref{fig:LSR}) is
531: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
532: \begin{equation}
533: -S_{I,I-1}^{\bf k_{\parallel }}\mathbf{C}_{I-1}+\left( P_{I,I}(\varepsilon
534: )-S_{I,I}^{\bf k_{\parallel }}\right) \mathbf{C}_{I}-S_{I,I+1}^{\bf k_{\parallel }}%
535: \mathbf{C}_{I+1}=0,  \label{eq:PSEoM}
536: \end{equation}%
537: where $\mathbf{C}_{I}\equiv C_{Ii}\equiv C_{IRlm}$ is a $(l_{\max }+1)^{2}H$
538: $\equiv M$ dimensional vector describing the amplitudes of the $I$-th layer
539: with $H$ sites and $(l_{\max }+1)^{2}$ orbitals per site. $P_{I,I}$
540: and $S_{I,J}$ are $M\times M$ matrices. $P_{I,I}$ is a diagonal matrix of
541: potential functions characterizing the AS potentials of layer $I$ and
542: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
543: \begin{equation}
544: S_{I,J}^{\bf k_{\parallel }}
545:      = \sum_{ {\bf T} \in  \{ {\bf T}_{I,J}\}}
546:            S^{\beta }  ({\bf T})
547:                e^{i {\bf k_{\parallel }.T}},
548: \end{equation}%
549: where $\left\{ {\bf T}_{I,J}\right\} $ denotes the set of vectors that
550: connect one lattice site in the $I$-th layer with all lattice sites in
551: the $J$-th layer.
552: 
553: By analogy with (\ref{eq:EoM}), equation~(\ref{eq:PSEoM}) is the
554: equation of motion we will use to calculate the amplitudes of right-
555: and left-going waves which determine the scattering matrix. We will
556: solve it for a fixed value of $\varepsilon $ (usually $\varepsilon _{F}$),
557: and some $\mathbf{k}_{\parallel }$ to find
558: $k_{\mu}(\varepsilon ,\mathbf{k}_{\parallel })$ the
559: component of the Bloch vector in the transport direction.
560: To keep the notation simple, explicit reference to the ${\bf k_{\parallel }}$
561: and $\varepsilon $ dependence will be omitted from now on.
562: The formalism to be described in the following can be applied to any
563: electronic structure code based on the KKR equation~(\ref{eq:PSEoM}),
564: such as third-generation TB-LMTO.\cite{Andersen:prb00,Tank:pssb00,Andersen:00}
565: 
566: 
567: Let us first consider the Bloch states in the ideal lead. To obtain linearly
568: independent solutions, we set $\mathbf{C}_{I}=\lambda ^{I}\mathbf{C}_{0}$,
569: since in a periodic potential the wave function should satisfy Bloch's
570: theorem. The potential function matrix is the same for all unit cells. The
571: structure constant matrix depends only on the relative positions and,
572: because that is how they are defined, there is only coupling between
573: adjacent principal layers so the equation of motion becomes
574: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
575: \begin{equation}
576: \left(
577: \begin{array}{cc}
578: S_{0,1}^{-1}(P-S_{0,0}) & -S_{0,1}^{-1}S_{1,0} \\
579: 1 & 0%
580: \end{array}%
581: \right) \left(
582: \begin{array}{l}
583: \mathbf{C}_{I} \\
584: \mathbf{C}_{I-1}%
585: \end{array}%
586: \right) =\lambda \left(
587: \begin{array}{l}
588: \mathbf{C}_{I} \\
589: \mathbf{C}_{I-1}%
590: \end{array}%
591: \right) ,
592: \label{eq:leads}
593: \end{equation}%
594: The eigenvalue $\lambda _{\mu }$ can be written in the form
595: $\lambda_{\mu}=exp(i\mathbf{k}_{\mu} \cdot\mathbf{T}_0)$
596: with $\mathbf{T}_0$ connecting equivalent sites in adjacent
597: prinicpal layers.
598: The wave vector ${\bf k}_{\mu }$ can be decomposed into
599: ${\bf k_{\parallel }}$ and a remainder which is in general not real,
600: ${\bf k}_{\mu } =
601: ({\bf k_{\parallel }},{\bf k}_{\mu }-{\bf k_{\parallel }})$.
602: Equation~(\ref{eq:leads}) has $2M$
603: eigenvalues and $2M$ eigenvectors, corresponding to $M$ right-going and $M$
604: left-going waves. By calculating the wavevectors and velocities  (see Eq.~\eqref{eq:velocity} 
605: and Appendix~\ref{sec:velocities}) of the lead
606: eigenmodes, the propagating and evanescent states can be identified and
607: sorted into right-going or left-going modes.
608: 
609: Letting $\mathbf{u}_{1}(-),...,\mathbf{u}_{M}(-)$ denote the left-going
610: solutions $\mathbf{C}_{0}$ corresponding to eigenvalues $\lambda
611: _{1}(-),...,\lambda _{M}(-)$ and $\mathbf{u}_{1}(+),...,\mathbf{u}_{M}(+)$
612: the right-going solutions corresponding to eigenvalues $\lambda
613: _{1}(+),...,\lambda _{M}(+)$, the matrix $U_{i\mu }(\pm )$ is defined as
614: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
615: \begin{equation}
616: U(\pm )=(\mathbf{u}_{1}(\pm )...\mathbf{u}_{M}(\pm ))  \label{eq:bigU}
617: \end{equation}%
618: and the matrix $\mathbf{\Lambda }(\pm )$ as the diagonal matrix with
619: elements $\lambda _{\mu }(\pm )$. Following Ando, we next expand any
620: left- or right-going wave, at $I=0$ for example, as
621: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
622: \begin{equation}
623: \mathbf{C}_{0}(\pm )=U(\pm )\mathbf{C}(\pm ).  \label{eq:bigC}
624: \end{equation}%
625: Note that $\mathbf{C}_{0}$ is a vector whose elements are labelled $i$
626: while the elements of the vector $\mathbf{C}$ are labelled $\mu $.%
627: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
628: \begin{equation}
629: F(\pm )\equiv U(\pm )\mathbf{\Lambda }_{\pm }U^{-1}(\pm )  \label{eq:bigF}
630: \end{equation}
631: is the matrix of Bloch factors (including evanescent states) transformed
632: onto the basis $\left\vert i\right\rangle $ and plays a central role in the
633: following. Knowing it makes it possible to translate a state expressed in
634: the basis $\left\vert i\right\rangle $ from layer $J$ of the lead to layer
635: $I$ by
636: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
637: \begin{equation}
638: \mathbf{C}_{I}(\pm )=F^{I-J}(\pm )\mathbf{C}_{J}(\pm ).  \label{eq:trans}
639: \end{equation}
640: 
641: \subsection{Scattering problem}
642: \label{ssec:scatt}
643: 
644: The scattering region $\mathcal{S}$, divided into $N$ principal layers
645: numbered $1$ to $N$, is now inserted between the left and right leads. The
646: resulting (scattering region + leads) problem is infinite dimensional in the
647: real space MTO representation but by making use of their translational
648: symmetry, the leads can be incorporated as boundary conditions and the
649: scattering problem can be reduced to a finite problem whose dimension is
650: determined by the size of the scattering region (number of sites $\times $
651: number of orbitals per site).
652: 
653: We set about decoupling the scattering region from the leads, first on the
654: left-hand side, then on the right. The amplitude in the $0$-th layer is
655: first separated into right- and left- going components $\mathbf{C}_{0}=%
656: \mathbf{C}_{0}(+)+\mathbf{C}_{0}(-)$. Because there is no additional
657: scattering in the leads, the right- and left-going components can be
658: translated to the left by one (principal layer) lattice spacing using the
659: generalized Bloch factors (\ref{eq:trans}) so the amplitude in layer $-1$
660: can be related to that in layer $0$ as
661: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
662: \begin{gather}
663: \mathbf{C}_{-1}=F_{\mathcal{L}}^{-1}(+)\mathbf{C}_{0}(+)+F_{\mathcal{L}%
664: }^{-1}(-)\mathbf{C}_{0}(-)  \notag \\
665: =\left[ F_{\mathcal{L}}^{-1}(+)-F_{\mathcal{L}}^{-1}(-)\right] \mathbf{C}%
666: _{0}(+)+F_{\mathcal{L}}^{-1}(-)\mathbf{C}_{0}.  \label{eq:C-1}
667: \end{gather}%
668: allowing us to express $\mathbf{C}_{-1}$ in terms of $\mathbf{C}_{0}$ and $%
669: \mathbf{C}_{0}(+)$ and so eliminate it from the equation of motion for the $%
670: 0 $-th layer
671: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
672: \begin{equation}
673: -S_{0,-1}\mathbf{C}_{-1}+\left( P_{0,0}-S_{0,0}\right) \mathbf{C}_{0}-S_{0,1}%
674: \mathbf{C}_{1}=0,
675: \end{equation}
676: which becomes
677: \begin{eqnarray}
678: &&(P_{0,0}-\widetilde{S}_{0,0})\mathbf{C}_{0}-S_{0,1}\mathbf{C}_{1}  \notag
679: \\
680: &=&S_{0,-1}\left[ F_{\mathcal{L}}^{-1}(+)-F_{\mathcal{L}}^{-1}(-)\right]
681: \mathbf{C}_{0}(+).
682: \end{eqnarray}
683: Here $\mathcal{L}$ denotes the left lead and
684: $\widetilde{S}_{0,0}=S_{0,0}+S_{0,-1}F_{\mathcal{L}}^{-1}(-)$.
685: $-S_{0,-1}F_{\mathcal{L}}^{-1}(-)$ is the ``embedding
686: potential'' for the left lead and the net result is that the equations
687: of motion have been truncated at layer $0.$
688: 
689: On the right-hand side of the scattering region, we are interested in the
690: situation where only right-going waves can exist in the $\left( N+1\right) $%
691: -th layer, so
692: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
693: \begin{equation}
694: \mathbf{C}_{N+2}=F_{\mathcal{R}}(+)\mathbf{C}_{N+1}(+)  \label{eq:CN+2}
695: \end{equation}
696: allowing us to eliminate $\mathbf{C}_{N+2}$\textbf{\ }from the EoM for
697: $\mathbf{C}_{N+1}$
698: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
699: \begin{equation}
700: (P_{N+1,N+1}-\widetilde{S}_{N+1,N+1})\mathbf{C}_{N+1}-S_{N+1,N}\mathbf{C}%
701: _{N}=0,  \label{hamnth}
702: \end{equation}%
703: where $\widetilde{S}_{N+1,N+1}=S_{N+1,N+1}+S_{N+1,N+2}F_{\mathcal{R}}(+)$
704: and $-S_{N+1,N+2}F_{\mathcal{R}}(+)$ is the embedding potential for the
705: right lead.
706: 
707: \begin{widetext}
708: 
709: Making use of the lead boundary conditions, the tail cancellation condition
710: for the scattering problem in real space is given by the set of
711: inhomogeneous linear equations
712: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
713: \begin{eqnarray}
714: \left( \begin{array}{cccccc}
715: (P -\widetilde{S})_{0,0} & -S_{0,1}    & 0           & \cdots & 0           & 0 \\
716: -S_{1,0}                 & (P-S)_{1,1} & -S_{1,2}    & \cdots & 0           & 0 \\
717: 0                        & -S_{2,1}    & (P-S)_{2,2} & \cdots & \vdots      & 0 \\
718: \vdots                   & \vdots      & \cdots      & \ddots & \vdots      & 0 \\
719: 0                        & 0           & \cdots      & \cdots & (P-S)_{N,N} & -S_{N,N+1} \\
720: 0                        & 0           & 0           & \cdots & -S_{N+1,N}  & (P-\widetilde{S})_{N+1,N+1}
721: \end{array} \right)
722: \left( \begin{array}{l}
723: \mathbf{C}_{0} \\
724: \mathbf{C}_{1} \\
725: \mathbf{C}_{2} \\
726: \vdots \\
727: \mathbf{C}_{N} \\
728: \mathbf{C}_{N+1}
729: \end{array} \right) & & \notag \\
730: \equiv
731: \left( \mathbf{P-}\widetilde{\mathbf{S}}\right)
732: \left(
733: \begin{array}{l}
734: \mathbf{C}_{0} \\
735: \mathbf{C}_{1} \\
736: \mathbf{C}_{2} \\
737: \vdots \\
738: \mathbf{C}_{N} \\
739: \mathbf{C}_{N+1}
740: \end{array}
741: \right)
742: = \left(
743: \begin{array}{c}
744: S_{0,-1}\left[ F_{\mathcal{L}}^{-1}(+)-F_{\mathcal{L}}^{-1}(-)\right] \mathbf{%
745: C}_{0}(+) \\
746: 0 \\
747: 0 \\
748: \vdots \\
749: 0 \\
750: 0%
751: \end{array}
752: \right) & &  \label{eq:green}
753: \end{eqnarray}%
754: which can be solved in terms of
755: $\mathbf{g=}\left( \mathbf{P-}\widetilde{\mathbf{S}}\right) ^{-1}$
756: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
757: \begin{equation*}
758: \left(
759: \begin{array}{l}
760: \mathbf{C}_{0} \\
761: \mathbf{C}_{1} \\
762: \mathbf{C}_{2} \\
763: \vdots \\
764: \mathbf{C}_{N} \\
765: \mathbf{C}_{N+1}%
766: \end{array}%
767: \right) =\mathbf{g}\left(
768: \begin{array}{c}
769: S_{0,-1}\left[ F_{\mathcal{L}}^{-1}(+)-F_{\mathcal{L}}^{-1}(-)\right] \mathbf{%
770: C}_{0}(+) \\
771: 0 \\
772: 0 \\
773: \vdots \\
774: 0 \\
775: 0%
776: \end{array}%
777: \right)
778: \end{equation*}
779: %
780: This treatment is very similar to the widely used surface Green function
781: method.\cite{Khomyakov:prb05} The boundary conditions in (\ref{eq:green})
782: are explicitly defined by considering the Bloch wave coming from the
783: left-hand side while for conventional retarded or advanced Green functions
784: the boundary conditions are specified by an infinitesimal imaginary part
785: of the energy parameter $\varepsilon $.
786: 
787: We are now in a position where we can relate the outgoing wave amplitude
788: in the right electrode to the incoming wave in the left electrode through
789: the Green function by
790: \begin{equation}
791: \mathbf{C}_{N+1}(+) =
792: \mathbf{C}_{N+1} =
793: g_{N+1,0}S_{0,-1}
794: \left[ F_{\mathcal{L}}^{-1}(+)-F_{\mathcal{L}}^{-1}(-)\right]
795: \mathbf{C}_{0}(+).
796: \end{equation}%
797: Using the transformation between the eigenstates and the localized basis
798: functions $U_{i\alpha }(\pm )$, we obtain the transmission and reflection
799: matrix elements\cite{Ando:prb91}
800: %%%%%%%%%%%%%%%%%%%%%%%%%
801: \begin{equation}
802: t_{\mu \nu }=\left( \frac{\upsilon _{\mu }}{\upsilon _{\nu }}\right)^{1/2}
803: \left\{ U_{\mathcal{R}}^{-1}(+)g_{N+1,0}S_{0,-1}
804: \left[ F_{\mathcal{L}}^{-1}(+)-F_{\mathcal{L}}^{-1}(-)
805: \right] U_{\mathcal{L}}(+)\right\}_{\mu \nu }
806: \label{trale}
807: \end{equation}
808: %%%%%%%%%%%%%%%%%%%%%%%%%
809: \begin{equation}
810: r_{\mu \nu }=\left( \frac{\upsilon _{\mu }}{\upsilon _{\nu }}\right)^{1/2}
811: \left\{ U_{\mathcal{L}}^{-1}(-)
812: \left\langle g_{0,0}S_{0,-1}
813: \left[ F_{\mathcal{L}}^{-1}(+)-F_{\mathcal{L}}^{-1}(-) \right] - 1\right\rangle
814: U_{\mathcal{L}}(+)\right\}_{\mu \nu },
815: \label{refle}
816: \end{equation}
817: %%%%%%%%%%%%%%%%%%%%%%%%%
818: where $\mu$ and $\nu $ label Bloch states and $\upsilon_{\mu }$,
819: $\upsilon_{\mu }$ are the components of the corresponding group
820: velocities in the transport direction.
821: Similarly, an incident wave from the right side is transmitted or
822: reflected as
823: %%%%%%%%%%%%%%%%%%%%%%%%%
824: \begin{equation}
825: t_{\mu \nu }^{\prime }=
826: \left( \frac{\upsilon _{\mu }}{\upsilon _{\nu }}\right) ^{1/2}
827: \left\{ U_{\mathcal{L}}^{-1}(-)g_{0,N+1}S_{N+1,N+2}
828: \left[ F_{\mathcal{R}}(-)-F_{\mathcal{R}}(+)\right] U_{\mathcal{R}}(-)
829: \right\}_{\mu\nu }
830: \label{trari}
831: \end{equation}
832: %%%%%%%%%%%%%%%%%%%%%%%%%
833: \begin{equation}
834: r_{\mu \nu }^{\prime }=
835: \left( \frac{\upsilon _{\mu }}{\upsilon _{\nu }}\right) ^{1/2}
836: \left\{ U_{\mathcal{R}}^{-1}(+)
837: \left\langle
838: g_{N+1,N+1}S_{N+1,N+2}\left[ F_{\mathcal{R}}(-)-F_{\mathcal{R}}(+)\right] -1
839: \right\rangle
840: U_{\mathcal{R}}(-) \right\} _{\mu \nu }.
841: \label{refri}
842: \end{equation}
843: The group velocities in (\ref{trale})-(\ref{refri}) are determined using the
844: expression
845: \begin{equation}
846:   \upsilon_{\mu }(\pm)=\frac{id}{\hbar}\left[
847:     \mathbf{u}^{\dagger}_{\mu}(\pm)S_{I,I+1}^{\mathbf{k}_{\parallel}}
848:     \mathbf{u}_{\mu}(\pm)\lambda_{\mu}-\mathrm{h.c.}
849:   \right]
850:   \label{eq:velocity}
851: \end{equation}
852: which is derived in Appendix A. Here, $I$ and $I+1$ denote
853: neighbouring principal layers in either left or right lead, 
854: $d= {\bf T}_0 \cdot \hat{n}$ is the distance between equivalent monolayers in
855: adjacent principal layers and $\hat{n}$ is a unit vector in the
856: transport direction.
857: \end{widetext}
858: 
859: \subsection{Disorder}
860: \label{ssec:Supercells}
861: 
862: Interfaces between materials with different lattice parameters
863: \cite{Xia:prl02} and disordered interfaces\cite{Xia:prb01,Zwierzycki:prb03}
864: can be modelled very flexibly using lateral supercells.
865: This approach allows us to study the effect of various types of disorder
866: on transport properties, ranging from homogeneous interdiffusion (alloying)
867: to islands, steps etc.
868: The supercell description of disorder becomes formally exact in the
869: limit of infinitely large supercells. In practice, satisfactory
870: convergence is achieved for supercells of quite moderate size
871: (see Sec. \ref{ssec:IntDis}).
872: 
873: 
874: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% IID XDISORDER: LEADS
875: \subsubsection*{Leads}
876: The factor limiting the ``size'' of scattering problem which can be handled
877: in practice is the rank of the 
878: blocks of the block-tridiagonal
879: scattering matrix in \eqref{eq:PSEoM}, which is proportional to the number
880: of atoms in the supercell.
881: If performed straightforwardly in the manner outlined in Sec.~\ref{ssec:Leads},
882: the solution of the lead equation \eqref{eq:leads} involves solving a
883: non-Hermitian eigenvalue problem whose rank is twice as large.
884: Unless use is made of the greater translational symmetry present in the
885: leads, this can become the limiting step in the whole calculation.
886: Doing so makes it possible to reduce the dimension of the lead state
887: calculation to a size determined by the dimension of a primitive unit
888: cell which is usually negligible.
889: 
890: \begin{figure}[bp]
891: \includegraphics[scale=0.60,clip=true]{supercell3.eps}
892: \caption{Illustration of lateral supercells and corresponding 2D
893: interface Brillouin zones. Top panel: lattice vectors for a
894: primitive unit cell containing a single atom (lhs) and a $4 \times
895: 4$ supercell (rhs). Bottom panel: a single k-point in the BZ (rhs)
896: corresponding to the $4 \times 4$ real-space supercell is
897: equivalent to $4 \times 4$ k-points in the BZ (lhs) corresponding
898: to the real-space primitive unit cell.} \label{fig:supercells}
899: \end{figure}
900: 
901: We consider an $H_1 \times H_2$ lateral supercell defined by the real-space
902: lattice vectors
903: \begin{equation}
904:   \mathbf{A}_1=H_1 \mathbf{a}_1
905:   \;\;\; \mathrm{and}\;\;\;
906:   \mathbf{A}_2=H_2 \mathbf{a}_2
907:   \label{eq:super_bigA}
908: \end{equation}
909: where $\mathbf{a}_1$ and $\mathbf{a}_2$ are  the lattice vectors describing
910: the in-plane periodicity of a primitive unit cell (Fig.~\ref{fig:supercells}).
911: The cells contained within the supercell are generated by the set of
912: translations
913: %
914: \begin{align}
915:   {\mathbf{T}_{\parallel}}_{\mathcal{T}} \in \mathbb{T}=&\left\{
916:       \mathbf{T}_{\parallel}=  h_1\mathbf{a}_1+ h_2\mathbf{a}_2
917:     \right. ; \nonumber\\
918:       & \left. 0 \le h_1 < H_1,\: 0 \le h_2 < H_2
919:     \right\}
920:   \label{eq:super_tau}
921: \end{align}
922: where $\mathcal{T}=1, \ldots , H_1\times H_2$ is a convenient cell index.
923: In reciprocal space the supercell Brillouin zone is defined by the
924: reduced vectors
925: %
926: \begin{equation}
927:  \mathbf{B}_1= \mathbf{b}_1 / H_1
928:   \;\;\; \mathrm{and}\;\;\;
929:   \mathbf{B}_2= \mathbf{b}_2 / H_2
930:   \label{eq:super_bigB}
931: \end{equation}
932: %
933: where $\mathbf{b}_1$ and $\mathbf{b}_2$ are the reciprocal lattice
934: vectors corresponding to the real space primitive unit cell. As a
935: result the Brillouin zone (BZ) is folded down, as shown schematically
936: in Fig.~\ref{fig:supercells} (bottom rhs), and the single
937: $\mathbf{k}^{\mathbb{S}}_{\parallel}$ point ($\mathbb{S}$ is used to
938: label supercell quantities) in the supercell BZ corresponds to the set
939: of $H_1 \times H_2$ ${\bf k}_{||}$ points in the original unfolded BZ
940: \begin{align}
941:   {\mathbf{k}_{\parallel}}_{\mathcal{K}} \in \mathbb{K}=&\left\{
942:       \mathbf{k}_{\parallel}= \mathbf{k}^{\mathbb{S}}_{\parallel}+ 
943:       h_1\mathbf{B}_1+ h_2\mathbf{B}_2
944:       \right. ; \nonumber\\
945:       & \left. 0 \le h_1 < H_1,\: 0 \le h_2 < H_2
946:     \right\}
947:   \label{eq:super_k}
948: \end{align}
949: %
950: with $\mathcal{K}=1, \ldots , H_1 \times H_2$.
951: Solutions associated with different ${\mathbf{k}_{\parallel}}_{\mathcal{K}}$
952: in the primitive unit cell representaton become different ``bands'' at the
953: single $\mathbf{k}^{\mathbb{S}}_{\parallel}$ in the supercell representaton.
954: 
955: The indices $\mathcal{T}$ and $\mathcal{K}$ provide a natural means of
956: describing the supercell-related matrices $U^\mathbb{S}(\pm)$ and
957: $F^\mathbb{S}(\pm)$ and their inverses in terms of $(H_1\times H_2)^2$
958: sub-blocks with dimensions defined by the primitive unit cell.
959: Thus $U^\mathbb{S}_{\mathcal{T} \mathcal{K}}(\pm)$ is the block
960: containing the amplitudes of the modes associated with
961: ${\mathbf{k}_{\parallel}}_{\mathcal{K}}$ in the
962: $\mathcal{T}$-th real-space cell.
963: 
964: Solving the single unit cell problem for the set of
965: $\mathbf{k}_{\parallel}$-points belonging to $\mathbb{K}$
966: (lhs of Fig.~\ref{fig:supercells}) and using the
967: Bloch symmetry of the eigenmodes, we get trivially
968: \begin{equation}
969:  U^\mathbb{S}_{\mathcal{T} \mathcal{K}}(\mathbf{k}^\mathbb{S}_{\parallel})
970:     =e^{i {\mathbf{k}_{\parallel}}_{\mathcal{K}}    \cdot
971:           {\mathbf{T}_{\parallel}}_{\mathcal{T}}           }
972:     U({\mathbf{k}_{\parallel}}_{\mathcal{K}})
973:   \label{eq:super_U}
974: \end{equation}
975: where $U({\mathbf{k}_{\parallel}}_{\mathcal{K}})$ is the matrix
976: \eqref{eq:bigU} of modes for a primitive unit cell for
977: ${\mathbf{k}_{\parallel}}_{\mathcal{K}}$ and the $\pm$ qualifier has
978: been dropped for simplicity. Defining the matrix of phase factors
979: \begin{equation}
980: X(\mathbf{k}^\mathbb{S}_{\parallel})=\left(
981:  \begin{array}{ccc}
982:    e^{i  {\mathbf{k}_{\parallel}}_1  \cdot {\mathbf{T}_{\parallel}}_1 } & \ldots &
983:    e^{i  {\mathbf{k}_{\parallel}}_H  \cdot {\mathbf{T}_{\parallel}}_1 }           \\
984:        \vdots &  & \vdots \\
985:                           \\
986:    e^{i  {\mathbf{k}_{\parallel}}_1  \cdot {\mathbf{T}_{\parallel}}_H } & \ldots  &
987:    e^{i  {\mathbf{k}_{\parallel}}_H  \cdot {\mathbf{T}_{\parallel}}_H }
988:     \end{array}
989:     \right)
990:   \label{eq:super_X}
991: \end{equation}
992: with $H \equiv H_1 \times H_2$, and its inverse $Y=X^{-1}$, we can
993: straightforwardly determine
994: \begin{equation}
995:   \left[
996:     U^{\mathbb{S}}(\mathbf{k}^\mathbb{S}_{\parallel})
997:   \right]^{-1}_{\mathcal{K} \mathcal{T}} =
998:   U^{-1}({\mathbf{k}_{\parallel}}_{\mathcal{K}})
999:                                Y_{\mathcal{K} \mathcal{T}}
1000:   \label{eq:super_Uinv}
1001: \end{equation}
1002: and
1003: \begin{equation}
1004:   F^\mathbb{S}_{\mathcal{T}_1 \mathcal{T}_2}
1005:                     (\mathbf{k}^\mathbb{S}_{\parallel})=
1006:   \sum_{\mathcal{K}}
1007:   X_{\mathcal{T}_1 \mathcal{K}} F({\mathbf{k}_{\parallel}}_{\mathcal{K}})
1008:              Y_{\mathcal{K} \mathcal{T}_2}
1009:   \label{eq:super_F}
1010: \end{equation}
1011: 
1012: The procedure outlined above for determining the matrices describing the
1013: lead modes scales linearly with the size of the supercell \emph{i.e.}, as
1014: ($H_1 \times H_2$) rather than as $(H_1\times H_2)^3$ which is the scaling
1015: typical for matrix operations. Another advantage is that it enables us to
1016: analyse the scattering. By keeping track of the relation between supercell
1017: ``bands'' and equivalent eigenmodes at different
1018: ${\mathbf{k}_{\parallel}}_{\mathcal{K}}$ (Fig.~\ref{fig:supercells}) we can
1019: straightforwardly obtain from \eqref{trale}-\eqref{refri}
1020: $t_{\mu \nu}   ({\mathbf{k}_{\parallel}}_{\mathcal{K}_1},
1021:                 {\mathbf{k}_{\parallel}}_{\mathcal{K}_2})$
1022: and other scattering coefficients. In other words the ``interband''
1023: specular scattering in the supercell picture translates, in the
1024: presence of disorder in the scattering region, into the ``diffuse''
1025: scattering between the $\mathbf{k}_{\parallel}$ vectors belonging
1026: to the $\mathbb{K}$ set.
1027: 
1028: \section{CALCULATIONS}
1029: \label{sec:Calc}
1030: 
1031: Even though the theoretical scheme outlined above contains no adjustable
1032: parameters, its practical implementation does involve numerous
1033: approximations, some physical, others numerical, which need to be evaluated.
1034: At present, any workable scheme must be based upon an independent particle
1035: approximation. The results of a transport calculation will be limited by
1036: the extent to which the single particle electronic structures used are
1037: consistent with the corresponding Fermi surfaces determined experimentally
1038: using methods such as de Haas-van Alphen measurements or the occupied and
1039: unoccupied electronic states close to the Fermi energy determined by, for
1040: example, photoelectron spectroscopy.
1041: 
1042: In this section we examine how various approximations affect our end
1043: results. We begin with the calculation of the scattering states in
1044: bulk Cu and bulk Co (\ref{ssec:CalcsLeads}).
1045: These are then used to study specular scattering from an ideal ordered
1046: Cu/Co(111) interface (\ref{ssec:OrdInt}) after which we describe how we
1047: model disordered interfaces (\ref{ssec:IntDis}) and how the results can
1048: be analysed.
1049: 
1050: \subsection{Leads}
1051: \label{ssec:CalcsLeads}
1052: 
1053: For a crystalline conductor with Bloch translational symmetry, each
1054: state at the Fermi energy can move unhindered through the solid so
1055: that the transmission matrix is diagonal with
1056: $\left\vert t_{\mu \nu }\right\vert ^{2}=\delta _{\mu \nu }$.
1057: In this \textit{ballistic} regime, (\ref{eq:LB}) reduces to
1058: \begin{equation}
1059: G^{\sigma }(\hat{n})
1060:        =\frac{e^{2}}{h}
1061:                 \sum_{\mu k_{\parallel }}
1062:                      |t_{\mu \mu}^{\sigma }({\bf k}_{\parallel })|^{2}
1063:        =\frac{e^{2}}{h} N^{\sigma }(\hat{n}).
1064:        \label{eq:Sharvin}
1065: \end{equation}
1066: and calculation of the so-called \textit{Sharvin} conductance becomes
1067: a matter of counting the number of modes (channels) propagating in the
1068: transport direction $\hat{n}$, denoted in \eqref{eq:Sharvin} as
1069: $N^{\sigma }(\hat{n})$. To solve \eqref{eq:leads} in practice, the
1070: orbital angular momentum expansions in (\ref{eq:TailExp}) and
1071: (\ref{eq:LCMTO}), which are in-principle infinite, must be truncated
1072: by introducing some cutoff in $l$, denoted $l_{\max }$.  Usually, a
1073: value of $l_{\max }=2$ or $3$ is used, corresponding to \textit{spd}-
1074: or \textit{spdf}-bases.
1075: 
1076: \begin{figure}[tbp]
1077: \includegraphics[scale=0.35,clip=true]{sharv_ins2_3.614.eps}
1078: \caption{Sharvin conductance $G^{\sigma }(111)$ (in units of
1079: $10^{15} \: \Omega^{-1} \mathrm{m}^{-2}$) for bulk {\em fcc} Cu
1080: and Co (majority and minority spin) plotted as a function of the
1081: normalized area element used in the Brillouin zone summation,
1082: $\Delta^2 {\bf k_{\parallel }} / A_{BZ} = 1/Q^2$. $Q$, the number
1083: of intervals along the reciprocal lattice vector is indicated at
1084: the top of the figure. Squares represent the series ($Q = 20, 40,
1085: 80, 160, 320$) least-squares fitted by the dashed line; diamonds,
1086: the series ($Q = 22, 44, 88, 176, 352$) least-squares fitted by
1087: the dash-dotted line. The part of the curve for the Co minority
1088: spin case to the left of the vertical dotted line is shown on an
1089: expanded scale in the inset. An \textit{fcc} lattice constant of
1090: $a=3.614\mathring{A}$ and \textit{spd} basis were used together
1091: with von Barth-Hedin's exchange-correlation potential.}
1092: \label{fig:lead_conv}
1093: \end{figure}
1094: 
1095: The $\mathbf{k}_{\parallel }$ summation is carried out by sampling,
1096: on a regular mesh, the 2D Brillouin zone (BZ) defined by the (lateral)
1097: translational periodicity perpendicular to $\hat{n}$.
1098: The results of carrying out this BZ summation are shown in Fig.~\ref{fig:lead_conv}
1099: where $G^{\sigma }(\hat{n})$ is plotted as a function of
1100: $\Delta^{2}k_{\parallel }/A_{BZ}$, the normalized area element per
1101: $\mathbf{k}_{\parallel }$-point for bulk {\em fcc} Cu and for the
1102: majority and minority spins of bulk \textit{fcc} Co.
1103: When the 2D-BZ reciprocal lattice vectors are each divided into $Q$
1104: intervals, then $\Delta^{2}k_{\parallel } / A_{BZ} = 1/Q^2$.
1105: It can be seen that the Sharvin conductance is converged to about 1\% if
1106: $3600=60\times 60$ points are used in the complete 2D-BZ and to about 0.2\%
1107: for $102400=320\times 320$ sampling points.
1108: The worst case is for the minority spin of Co which has a complex
1109: multi-sheeted Fermi surface.
1110: To see if there are any simple underlying trends in the convergence, we
1111: repeatedly bisect the intervals used in the BZ summation starting with
1112: $Q=20$ and $Q=22$, shown in the figure as squares and diamonds, respectively
1113: and least-squares fitted with the dashed and dash-dotted lines.
1114: The convergence is fairly uniform but not very systematic indicating that
1115: the summation is limited by fine structure in the integrand at the smallest
1116: length scale studied which can only be resolved by increasingly fine sampling.
1117: Thus there is nothing to be gained by developing more sophisticated
1118: interpolation schemes and when we introduce disorder in
1119: Section~\ref{ssec:IntDis}, this will be even more so.
1120: However, in the following we will see that the level of convergence we can
1121: achieve with discrete sampling is quite adequate and not a limiting step
1122: in the whole procedure.
1123: 
1124: The calculations shown in the figure were performed using an
1125: $spd$-basis, for an \textit{fcc} lattice constant $a=3.614 \,
1126: \mathring{A}$ corresponding to the experimental volume of bulk
1127: (\textit{fcc}) Cu and using the exchange-correlation potential
1128: calculated and parameterized by von Barth and
1129: Hedin.\cite{vonBarth:jpc72} For convenience, and to avoid repetition,
1130: we will refer to this in the following as a ``standard''
1131: configuration.  The converged values are given (underlined) in
1132: Table~\ref{tab:A} together with values calculated using an
1133: \textit{fcc} lattice constant $a=3.549 \, \mathring{A}$ corresponding
1134: to the volume of bulk \textit{hcp} Co.\cite{fn:WS_radii} Because we
1135: shall be studying Cu/Co interfaces where the volume per atom is not
1136: known very precisely from experiment, we will want to estimate the
1137: variation that can be expected when different but equally reasonable
1138: lattice constants are used. The increase of 3.4\% (from 0.558 to
1139: $0.577 \times 10^{15} \: \Omega^{-1} \mathrm{m}^{-2}$) observed for Cu
1140: can be attributed to the increased areal density of Cu atoms,
1141: $(3.614/3.549)^2$ corresponding to $\sim 3.7\%$. The Table also
1142: contains the corresponding results obtained with an
1143: \textit{spdf}-basis.  To the numerical accuracy shown, there is no
1144: difference between the \textit{spd} and \textit{spdf} case for Cu.
1145: 
1146: For Co majority spin states, there is a 4\% decrease in the
1147: conductance on going from an \textit{spd} to an \textit{spdf} basis.
1148: For a lattice constant $a=3.614 \, \mathring{A}$, the magnetization is
1149: 1.684$\, \mu_B$/atom for an \textit{spd}- and
1150: 1.648$\, \mu_B$/atom for an \emph{spdf} basis corresponding, respectively,
1151: to $n_{\rm maj}=$ 5.342 and 5.324 electrons in the majority spin bands.
1152: Since all five (nominal) majority-spin $d$ bands are full there are
1153: 0.342 and 0.324 electrons in the free-electron-like $sp$ band. In a free
1154: electron
1155: picture the ratio of the projection of the spherical Fermi surfaces is
1156: $(0.324/0.342)^{2/3}=0.96$, thus explaining the observed numerical
1157: result.
1158: 
1159: \begin{table}[btp]
1160: \begin{ruledtabular}
1161: \begin{tabular}{lcllll}
1162:          &           &        &  & \multicolumn{2}{c} {$G^{\sigma }(111)$}\\ \cline{5-6}
1163:          & $a$(\AA ) & basis  & $n_{\sigma}$ & present calc. &
1164:                                                     Schep\footnotemark[1] \\ \hline
1165: Copper   & 3.549     & $spd$  & 5.5   & 0.577(0.577,0.577)       & 0.57 \\
1166:          & 3.549     & $spdf$ & 5.5   & 0.577(0.577)             & ---- \\
1167:          & 3.614     & $spd$  & 5.5   & \underline{0.558}(0.559) & 0.55 \\
1168:          & 3.614     & $spdf$ & 5.5   & 0.558(0.558)             & 0.55 \\
1169: \\
1170: Cobalt   & 3.549     & $spd$  & 5.323 & 0.469(0.459,0.467)       & 0.45 \\
1171: majority & 3.549     & $spdf$ & 5.304 & 0.449(0.440)             & 0.43 \\
1172:          & 3.614     & $spd$  & 5.342 & \underline{0.466}(0.457) & 0.45 \\
1173:          & 3.614     & $spdf$ & 5.324 & 0.448(0.439)             & ---- \\
1174: \\
1175: Cobalt   & 3.549     & $spd$  & 3.677 & 1.082(1.081,1.082)       & 1.10 \\
1176: minority & 3.549     & $spdf$ & 3.696 & 1.120(1.125)             & 1.13 \\
1177:          & 3.614     & $spd$  & 3.658 & \underline{1.046}(1.047) & 1.06 \\
1178:          & 3.614     & $spdf$ & 3.676 & 1.074(1.079)             & ---- \\
1179: 
1180: \end{tabular}
1181: \end{ruledtabular}
1182: \footnotetext[1]{Ref.\onlinecite{Schep:prb98} }
1183: \caption[Tab1]{
1184: The Sharvin conductances per spin (in units of $10^{15} \:
1185: \Omega^{-1} \mathrm{m}^{-2}$) in the (111) direction for
1186: \emph{fcc} Cu and Co using the experimental volumes of Cu and Co.
1187: The underlined numbers are the converged values discussed in relation
1188: to Fig.~\ref{fig:lead_conv}.
1189: Most of the results were obtained with von Barth-Hedin's
1190: exchange-correlation potential while the results in brackets are
1191: for Perdew-Zunger (PZ) and Vosko-Wilk-Nusair (VWN) parameterizations,
1192: respectively. Where a single number is given in brackets, it means
1193: that PZ and VWN potentials yield identical results to the accuracy
1194: given.
1195: The corresponding results of Schep {\em et al.} are
1196: given in the last column. The number of electrons with spin
1197: $\sigma$ is given in the fourth column. }
1198: \label{tab:A}
1199: \end{table}
1200: 
1201: The Co majority-spin conductance scarcely changes with changing
1202: lattice constant, however. The origin of this behaviour lies in the
1203: volume dependence of the magnetic moment. When the lattice constant is
1204: decreased, the $d$ bands broaden and the magnetic moment decreases from
1205: 1.684 to 1.646$\, \mu_B$/atom in the $spd$ case with a corresponding
1206: decrease of the occupancy of the $sp$ band from 0.342 to 0.323
1207: majority-spin electrons. The corresponding 4\% decrease in conductance
1208: is almost perfectly compensated by the increased areal density of atoms
1209: so there is no net change.
1210: For the minority-spin conductance, the same factors play a role but
1211: now the $d$ bands are only partly filled. This results in complex
1212: Fermi surfaces for which simple estimates cannot be made. In this
1213: case recourse must be made to full band structure calculations.
1214: We return to this in Sect.~\ref{ssec:OrdInt}.
1215: 
1216: The calculations presented so far were carried out using the
1217: exchange-correlation potential calculated and parameterized by
1218: von Barth and Hedin.\cite{vonBarth:jpc72}
1219: This is only one of a number of potentials we could have used, none of
1220: which is clearly better than the others in describing the ground state
1221: properties of magnetic materials.
1222: To gauge the uncertainty arising from this arbitrary choice, a number of
1223: calculations were carried out using the potentials given by
1224: Perdew-Zunger\cite{Perdew:prb81} and Vosko-Wilk-Nusair\cite{Vosko:cjp80}
1225: and the results are given in brackets in the Table. Using different
1226: exchange-correlation potentials leads to variation in the conductances
1227: of the order of 1 or 2 percent.
1228: 
1229: A different (but equivalent) approach was adopted by
1230: Schep \emph{et al.}\cite{Schep:prl95,Schep:prb98} to
1231: the determination of the Sharvin conductances for the same systems using
1232: conventional first-principles LMTO-ASA bulk electronic band structures,
1233: i.e. using $\varepsilon _{i}(\mathbf{k})$ rather than
1234: $k_{\mu}(\varepsilon =\varepsilon _{F},\mathbf{k}_{\parallel })$ as used
1235: here. He expressed
1236: the Sharvin conductance as a projection of the Fermi surface onto a plane
1237: perpendicular to the transport direction and calculated the areas using
1238: a suitably modified 3D-BZ integration scheme. His results are also given in
1239: Table~\ref{tab:A} and are as consistent with our present values as can be
1240: expected when using two entirely different computer codes.
1241: 
1242: In determining the conductance of the leads, the BZ summation does not
1243: present a problem. The uncertainties arising from small variations in
1244: the atomic volumes, from incompleteness of the basis and from the choice
1245: of LDA parameterization are of comparable size. The MTO-AS approximation
1246: can be systematically improved but only at substantial computational
1247: cost. Since there is currently no way to systematically improve upon the
1248: LDA we identify it and the lack of knowledge of the atomic structure
1249: as limiting factors in studying transport from first principles. Though
1250: the atomic structures could be determined theoretically by total energy
1251: minimization, the LDA again presents a barrier since it systematically
1252: underestimates lattice constants of transition metals in particular of
1253: the 3$d$ series. Gradient corrections sometimes yield improvements but
1254: unfortunately not systematically so. We conclude that our knowledge of
1255: and ability to calculate from first principles Fermi surfaces for bulk
1256: magnetic materials such as Fe or Co does not at present justify using
1257: a more accurate but substantially more expensive computational scheme
1258: than the present one.
1259: 
1260: \subsection{Ordered Interfaces}
1261: \label{ssec:OrdInt}
1262: 
1263: Cu and Co have slightly different atomic volumes. The equilibrium lattice
1264: constant of Cu is $3.614 \, \mathring{A}$ and of Co $3.549 \, \mathring{A}$,
1265: assuming an {\em fcc} structure. Even in the absence of interface disorder,
1266: the lattice spacing will not be homogeneous and will depend on the lattice
1267: constant of the substrate on which the sample was grown, on the global and
1268: local concentrations of Cu and Co, and on other details of how the structure
1269: was prepared. In principle we could calculate all of this by energy
1270: minimization. However, we judge that the additional effort needed is not
1271: justified by current experimental knowledge. Instead, we content
1272: ourselves with estimating the uncertainty which results from plausible
1273: variations in the (interface) structure by considering two limiting cases
1274: and one intermediate case. In each case an {\em fcc} structure is assumed,
1275: with lattice constants corresponding to
1276: (i) the atomic volume of Cu,
1277: (ii) the atomic volume of Co,
1278: (iii) an intermediate case with arithmetic mean of Cu and Co atomic volumes.
1279: \begin{table}[b]
1280: \begin{ruledtabular}
1281: \begin{tabular}{l D{.}{.}{4} D{.}{.}{5} D{.}{.}{5} D{.}{.}{5}}
1282: \multicolumn{1}{c}{$a (\mathring{A}$)} &
1283:        \multicolumn{2}{c} {3.549} &
1284:                          \multicolumn{1}{c} {3.582} &
1285:                                      \multicolumn{1}{c} {3.614} \\
1286: \hline
1287: \multicolumn{1}{c}{Basis} &
1288:     \multicolumn{1}{c}{\em spdf} &
1289:                       \multicolumn{1}{c}{\em spd} &
1290:                                     \multicolumn{1}{c}{\em spd} &
1291:                                                  \multicolumn{1}{c}{\em spd} \\
1292: \hline
1293: $m_{\rm Cu}$(bulk)  &  0.000      &  0.000        &  0.000      &  0.000     \\
1294: $m_{\rm Cu}$(int-4) &  0.001(1)   &  0.001        &  0.001      &  0.001     \\
1295: $m_{\rm Cu}$(int-3) & -0.001(0)   &  0.000        &  0.000      &  0.000     \\
1296: $m_{\rm Cu}$(int-2) & -0.005(5)   & -0.005(4,5)   & -0.005(4)   & -0.005     \\
1297: $m_{\rm Cu}$(int-1) &  0.002(4)   &  0.004(6,4)   &  0.003(4)   &  0.001(2)  \\
1298: \hline
1299: $m_{\rm Co}$(int+1) & 1.526(490)  &  1.578(45,73) &  1.605(573) &  1.636(01) \\
1300: $m_{\rm Co}$(int+2) & 1.621(597)  &  1.656(35,53) &  1.673(53)  &  1.690(70) \\
1301: $m_{\rm Co}$(int+3) & 1.602(576)  &  1.645(21,41) &  1.662(39)  &  1.680(59) \\
1302: $m_{\rm Co}$(int+4) & 1.610(587)  &  1.649(27,45) &  1.665(45)  &  1.683(62) \\
1303: $m_{\rm Co}$(bulk)  & 1.609(590)  &  1.646(22,42) &  1.667(45)  &  1.684(62) \\
1304: \hline
1305: $G^{\rm maj}(111)$  &  0.409(399) &  0.431(21,29) &  0.433(22)  &  0.434 (24) \\
1306: $G^{\rm min}(111)$  &  0.378(379) &  0.378(80,79) &  0.371(73)  &  0.364 (67)
1307: \end{tabular}
1308: \end{ruledtabular}
1309: \caption{Variation of the layer-resolved magnetic moments (in Bohr
1310: magnetons) for Cu/Co(111) interfaces with basis set and lattice constant.
1311: These results were obtained with von Barth-Hedin's exchange-correlation
1312: potential while the results in brackets, where given, are for
1313: Perdew-Zunger and Vosko-Wilk-Nusair parameterizations, respectively.
1314: In the last two rows, the interface conductances are given in units of
1315: $10^{15} \: \Omega^{-1} \mathrm{m}^{-2}$.}
1316: \label{tab:B}
1317: \end{table}
1318: 
1319: Our starting point is a self-consistent TB-LMTO SGF calculation\cite{Turek:97}
1320: for the interface embedded between semi-infinite Cu and Co leads whose
1321: potentials and spin-densities were determined self-consistently in separate
1322: ``bulk'' calculations. The charge and spin-densities are allowed to vary in
1323: $n_{\rm Cu}$ layers of Cu and $n_{\rm Co}$ layers of Co bounding the interface.
1324: The results of these calculations for Cu/Co(111) interfaces and the three
1325: different lattice constants detailed above are given in Table~\ref{tab:B}
1326: for $n_{\rm Cu}$=4, $n_{\rm Co}$=4. In the Cu layers, only tiny moments are
1327: induced. Only four layers away from the interface on the Co side, the
1328: magnetic moments are seen to be very close to the bulk values. At the
1329: interface, where the $d$-bandwidth is reduced as a result of the lower
1330: coordination number, the moments are {\em suppressed} rather than enhanced.
1331: This occurs because the majority-spin $d$ bands are full and their number
1332: cannot increase. The width of the free-electron like $sp$ band is less
1333: sensitive to the change in coordination and its exchange splitting also
1334: changes less. As a result, there is little change in the $sp$ moment.
1335: When the $d$-bandwidth is reduced, there is conversion of minority- and
1336: majority-spin $sp$ electrons, without loss of the $sp$ moment, to the
1337: minority-spin $d$ band with loss of $d$ moment. This picture is supported
1338: by the full calculations.
1339: 
1340: Earlier we saw that an $\sim 2\%$ change in lattice constant changed
1341: the bulk magnetic moment of \textit{fcc} Co by $2.3 \%$. The effect of
1342: changing the basis, from \textit{spd} to \textit{spdf}, was similar.
1343: From Table~\ref{tab:B}, the interface moments are seen to behave in a
1344: comparable fashion.
1345: The magnetic moment of the interface Co atoms decreases by $3.7 \%$,
1346: from 1.636$\, \mu_B$/atom for $a=3.614 \, \mathring{A}$
1347: to 1.578$\, \mu_B$/atom for $a=3.549 \, \mathring{A}$ for an $spd$ basis
1348: and decreases from 1.578$\, \mu_B$/atom to 1.526$\, \mu_B$/atom for
1349: an $spdf$ basis for $a=3.549 \, \mathring{A}$, a change of $3.4 \%$.
1350: Thus the $sp$ to $d_{\rm min}$ conversion is enhanced at the interface
1351: by the reduced $d$-bandwidth.
1352: 
1353: \begin{figure}[tbp]
1354: \includegraphics[scale=0.30,clip=true]{cond_3.614.eps}
1355: \caption{Interface conductance $G^{\sigma }(111)$ (in units of
1356: $10^{15} \: \Omega^{-1} \mathrm{m}^{-2}$) for an {\em fcc}
1357: Cu/Co(111) interface for majority and minority spins plotted as a
1358: function of the normalized area element used in the Brillouin zone
1359: summation, $\Delta^2 {\bf k_{\parallel }} / A_{BZ} = 1/Q^2$. $Q$,
1360: the number of intervals along the reciprocal lattice vector is
1361: indicated at the top of the figure. Squares represent the series
1362: ($Q = 20, 40, 80, 160, 320$) least-squares fitted by the dashed
1363: line; diamonds, the series ($Q = 22, 44, 88, 176, 352$)
1364: least-squares fitted by the dash-dotted line. The part of the
1365: curve for the Co minority spin case to the left of the vertical
1366: dotted line is shown on an expanded scale in the inset.
1367: A ``standard'' configuration was used.\cite{fn:standard}}
1368: \label{fig:cl_int_conv}
1369: \end{figure}
1370: 
1371: Once the interface potential has been obtained, the transmission matrix
1372: can be calculated and the BZ summation carried out. The convergence of
1373: this summation, shown in Fig.~\ref{fig:cl_int_conv} for a lattice
1374: constant of $a=3.614 \, \mathring{A}$ and an \textit{spd} basis closely
1375: parallels that seen in Fig.~\ref{fig:lead_conv} and therefore the
1376: k-summation does not represent a limitation in practice. Converged
1377: conductances
1378: \begin{equation}
1379: G^{\sigma }(\hat{n})
1380: = \frac{e^2}{h} \sum_{\mu,\nu,{\bf k_{\parallel}}}
1381:                 T_{\mu \nu}^{\sigma } ({\bf k_{\parallel }})
1382: = \frac{e^2}{h} \sum_{\mu,\nu,k_{\parallel}}
1383:                |t_{\mu \nu }^{\sigma }({\bf k_{\parallel }})|^2
1384: \label{eq:TransMat}
1385: \end{equation}
1386: are given in the last two rows of Table~\ref{tab:B}.
1387: Though we will not concern ourselves in this publication with the
1388: application of the formalism we have been developing to a detailed
1389: interpretation of experimental observations, it should be noted that
1390: even a modest spin-dependence of ``bare'' interface conductances
1391: ($\sim 20 \%$) can lead to spin-dependent interface resistances
1392: differing by a factor of $\sim 3-5$ once account is taken of the
1393: finiteness of the conductance of the perfect leads using a formula
1394: derived by Schep {\em et al.}\cite{Schep:prb97}
1395: \begin{equation}
1396: R_{A/B}^{\sigma }=\frac{h}{e^{2}}\left[
1397: \frac{1}{\sum T_{\mu \nu }^{\sigma }} -
1398: \frac{1}{2}\left(
1399: \frac{1}{N_{A}^{\sigma }}+\frac{1}{N_{B}^{\sigma }}\right) \right]
1400: \label{eq:R_Schep}
1401: \end{equation}
1402: where $N_A^{\sigma }$ and $N_B^{\sigma }$ are the Sharvin conductances
1403: (see Eq. \eqref{eq:Sharvin}) of the materials A and B forming the
1404: interface, in units of ${e^2 / h}$.
1405: 
1406: The majority-spin case can be readily understood in terms of the
1407: geometry of the Fermi surfaces of Cu and Co so we begin by discussing
1408: this simple case before examining the more complex minority-spin
1409: channel.
1410: 
1411: \subsubsection*{Clean Cu/Cu (111) Interface: Majority Spins}
1412: 
1413: In the absence of disorder, crystal momentum parallel to the interface
1414: is conserved. If, for a given value of ${\bf k}_{\parallel}$, there
1415: is a propagating state in Cu incident on the interface but none in Co,
1416: then an electron in such a state is completely reflected at the interface.
1417: Conversely, ${\bf k}_{\parallel}$'s for which there is a propagating state
1418: in Co but none in Cu also cannot contribute to the conductance.
1419: To determine the existence of such states, it is sufficient to inspect
1420: projections of the Fermi surfaces of {\em fcc} Cu and majority-spin
1421: Co onto a plane perpendicular to the transport direction ${\hat n}$,
1422: shown in Fig.~\ref{fig:CuCo111_maj} for ${\hat n}=(111)$. The first
1423: feature to note in the figure (left-hand and middle panels) is that per
1424: ${\bf k}_{\parallel}$ there is only a single channel with positive group
1425: velocity so that the transmission matrix in \eqref{eq:TransMat} is a
1426: complex number whose modulus squared is a transmission probability with
1427: values between 0 and 1. It is plotted in the
1428: right-hand panel and can be interpreted simply. Regions which are
1429: depicted blue correspond to ${\bf k}_{\parallel}$'s for which there are
1430: propagating states in Cu but none in Co. These states have transmission
1431: probability 0 and are totally reflected. For values of
1432: ${\bf k}_{\parallel}$ for which there are propagating states in both
1433: Cu and Co, the transmission probability is very close to one, depicted
1434: red. These states are essentially free electron-like states which have
1435: the same symmetry in both materials and see the interface effectively
1436: as a very low potential step.
1437: Close to the centre of the figure there is an annular region where
1438: there are propagating states in Co but none in Cu so they do not
1439: contribute to the conductance.
1440: Performing the sum in \eqref{eq:TransMat}, we arrive at an interface
1441: conductance of $0.434 \times 10^{15} \: \Omega^{-1} \mathrm{m}^{-2}$
1442: to be compared to the Sharvin conductances given in Table~\ref{tab:A}
1443: for Cu and Co; for $a=3.614 \, \mathring{A}$ and an $spd$ basis these
1444: are, respectively, 0.558 and 0.466 in the same units. The interface
1445: conductance of 0.434 is seen to be essentially the Sharvin conductance
1446: of the majority states of Co reduced because the states closest to the
1447: $\Lambda$-axis (corresponding to the symmetry axis of the figures, the
1448: $\Gamma L$ line in reciprocal space) do not contribute. The explanation
1449: of the 5\% decrease found on going from an $spd$ to an $spdf$ basis,
1450: (0.431 to 0.409), parallels that given for the corresponding change in
1451: the Sharvin conductance of bulk Co (0.469 to 0.449 in Table~\ref{tab:A}).
1452: 
1453: \begin{figure}[tbp]
1454: \includegraphics[scale=0.43,clip=true]{CuCo_111_maj_cm.eps}
1455: \caption{ Top row, left-hand panel: Fermi surface (FS) of Cu;
1456: middle panel: majority-spin FS of Co; right-hand panel: Cu FS
1457: viewed along the (111) direction with a projection of the bulk fcc
1458: Brillouin zone (BZ) onto a plane perpendicular to this direction
1459: and of the two dimensional BZ.
1460: %
1461: Bottom row, left-hand and middle panels: projections onto a plane
1462: perpendicular to the (111) direction of the Cu and majority-spin
1463: Co Fermi surfaces;
1464: right-hand panel: transmission probability for majority-spin states
1465: as a function of transverse crystal momentum, $T({\bf k_{\parallel}})$
1466: for an {\em fcc} Cu/Co(111) interface.
1467: A ``standard'' configuration was used.\cite{fn:standard}
1468: }
1469: \label{fig:CuCo111_maj}
1470: \end{figure}
1471: 
1472: \subsubsection*{Clean Cu/Cu (111) Interface: Minority Spins}
1473: 
1474: The minority-spin case is considerably more complex because the
1475: Co minority-spin $d$ bands are only partly filled, resulting in
1476: multiple sheets of Fermi surface. These sheets are shown in
1477: Fig.~\ref{fig:CuCo111_min} together with their projections onto a
1478: plane perpendicular to the (111) transport direction. Compared to
1479: Fig.~\ref{fig:CuCo111_maj}, one difference we immediately notice
1480: is that even single Fermi surface (FS) sheets are not single valued:
1481: for a given ${\bf k}_{\parallel}$ there can be more than one mode
1482: with positive group velocity. The areas depicted green in the
1483: projections of the FS sheets from the fourth and fifth bands are
1484: examples where this occurs.
1485: 
1486: \begin{figure*}[btp]
1487: \includegraphics[scale=0.95,clip=true]{CuCo_111_min_cm.eps}
1488: \caption{ Top row, lefthand panel: Fermi surface (FS) of
1489: \emph{fcc} Cu; middle panels: third, fourth and fifth FS sheets of
1490: minority-spin fcc Co; righthand panel: projection of the bulk fcc
1491: Brillouin zone (BZ) onto a plane perpendicular to the (111)
1492: direction and of the two dimensional BZ.
1493: %
1494: Middle row: corresponding projections of individual FS sheets and (rhs)
1495: of Co total. The number of propagating states with positive velocity is
1496: colour-coded following the colour bar on the right.
1497: %
1498: Bottom row: probability $T_{\mu \nu}(\mathbf{k_{\parallel }})$
1499: for a minority-spin state on the single FS sheet of Cu ($\nu = 1$)
1500: to be transmitted through a Cu$|$Co(111) interface into FS sheet
1501: $\mu$ of \emph{fcc} Co as a function of the transverse crystal
1502: momentum $\mathbf{k_{\parallel }}$. The point Y is such that there
1503: are only propagating states in Cu and in the fourth FS sheet of
1504: Co. For the point Y' slightly further away from $\Lambda $ and
1505: indicated by a small open square there is, in addition, a
1506: propagating state in the third FS sheet of Co.
1507: Results are for a ``standard'' configuration.\cite{fn:standard}
1508:  }
1509: \label{fig:CuCo111_min}
1510: \end{figure*}
1511: 
1512: An electron incident on the interface from the Cu side, with transverse
1513: crystal momentum
1514: ${\bf k}_{\parallel}$, is transmitted into a linear combination of
1515: all propagating states with the same ${\bf k}_{\parallel}$ in Co; the
1516: transmission matrix $t_{\mu \nu }^{\sigma }({\bf k_{\parallel }})$ is
1517: in general not square but rectangular. The transmission probabilities
1518: $T_{\mu \nu}({\bf k_{\parallel }})$ are shown in the bottom row of
1519: Fig.~\ref{fig:CuCo111_min}. Because there is only a single incident
1520: state for all ${\bf k}_{\parallel}$, the maximum transmission
1521: probability is one. Comparison of the total minority-spin
1522: transmission probability $T_{\mathcal{LR}}({\bf k_{\parallel}})$
1523: (Fig.~\ref{fig:CuCo111_min}, bottom right-hand panel) with the
1524: corresponding majority-spin quantity (right-hand panel of
1525: Fig.~\ref{fig:CuCo111_maj}) strikingly illustrates the spin-dependence
1526: of the interface scattering, much more so than the integrated
1527: quantities might have led us to expect; the interface conductances,
1528: 0.364 and $0.434 \times 10^{15} \: \Omega^{-1} \mathrm{m}^{-2}$
1529: from Table~\ref{tab:B}, differ by only $\sim 20\%$.
1530: 
1531: Three factors contribute to the large ${\bf k}_{\parallel}$-dependence
1532: of the transmission probability: first and foremost, the complexity of
1533: the Fermi surface of both materials but especially of the minority spin
1534: of Co; secondly and inextricably linked with the first because of the
1535: relationship $\hbar\upsilon_{\bf k}={\nabla}_k \varepsilon({\bf{k}})$,
1536: the mismatch of the Fermi velocities of the states on either side of
1537: the interface. Thirdly, the orbital character of the states $\mu$ and
1538: $\nu$ which varies strongly over the Fermi surface and gives rise to
1539: large matrix element effects.
1540: 
1541: The great complexity of transition metal Fermi surfaces, clear from the
1542: figure and well-documented in standard textbooks, is not amenable to
1543: simple analytical treatment and has more often than not been neglected
1544: in theoretical transport studies. Nevertheless, as illustrated
1545: particularly well by the ballistic limit,\cite{Schep:prl95,Schep:prb98}
1546: spin-dependent band structure effects have been shown to lead to
1547: magnetoresistance ratios comparable to what are observed experimentally
1548: in the current-perpendicular-to-plane (CPP) measuring configuration and
1549: cannot be simply ignored in any quantitative discussion.
1550: Most attempts to take into account contributions of the $d$ states to
1551: electronic transport do so by mapping the five $d$ bands onto a single
1552: tight-binding or free-electron band with a large effective mass.
1553: 
1554: Fermi surface topology alone cannot explain all aspects of the
1555: tranmission coefficients seen in Fig.~\ref{fig:CuCo111_min}. For
1556: example, there are values of ${\bf k_{\parallel}}$, such as that
1557: labelled $Y$ in the figure, for which propagating solutions exist on
1558: both sides of the interface yet the transmission probability is zero.
1559: This can be understood as follows. 
1560: At ${\bf k_{\parallel}} =Y$, the propagating states in Cu have \{$s,
1561: p_y, p_z, d_{yz}, d_{3z^2 - r^2}, d_{x^2 - y^2}$\} character (assuming
1562: the choice of in-plane axes as illustrated in the top righthand panel
1563: of Fig.~\ref{fig:CuCo111_min}) and are even with respect to
1564: reflection in the plane formed by the $y$-axis and the transport
1565: direction perpendicular to the (111) plane which we choose to be the
1566: $z$-axis.  For this ${\bf k_{\parallel}}$ the only propagating state
1567: in Co is in the fourth band. It has \{$p_x, d_{xy}, d_{xz}$\}
1568: character which is odd with respect to reflection in the $yz$ plane.
1569: Consequently, the corresponding hopping matrix elements in the
1570: Hamiltonian (and in the Green's function) vanish and the transmission
1571: is zero.
1572: 
1573: Along the $k_y$ axis the symmetry of the states in Cu and those in the
1574: fourth band of Co remain the same and the transmission is seen to vanish
1575: for all values of $k_y$.
1576: However, at points further away from $\Lambda$, we encounter states in
1577: the third band of Co which have even character whose matrix elements do
1578: not vanish by symmetry and we see substantial transmission probabilities.
1579: Similarly, for points closer to $\Lambda$, there are states in the
1580: fifth band of Co with even character whose matrix elements also do not
1581: vanish and again the transmission probability is substantial. Because
1582: it is obtained by superposition of transmission probabilities from Cu
1583: into the third, fourth and fifth sheets of the Co FS, the end result,
1584: though it may appear very complicated, can be straightforwardly
1585: analysed in this manner k-point by k-point.
1586: 
1587: Though the underlying lattice symmetry is only threefold, the Fermi
1588: surface projections shown in Fig.~\ref{fig:CuCo111_min} have six-fold
1589: rotational symmetry about the line $\Lambda$ because the bulk $fcc$
1590: structure has inversion symmetry (and time-reversal symmetry). The
1591: interface breaks the inversion symmetry so $T_{\mu \nu}({\bf
1592:   k_{\parallel }})$ has only threefold rotation symmetry for the
1593: individual FS sheets. However, in-plane inversion symmetry is
1594: recovered for the total transmission probability $T_{\mathcal{LR}}( -
1595: {\bf k_{\parallel}}) =T_{\mathcal{LR}}( {\bf k_{\parallel}})$ which
1596: has full sixfold symmetry. This follows from the time-reversal
1597: symmetry and is proven in Appendix~\ref{sec:symmetry}.
1598: 
1599: \subsection{Interface Disorder}
1600: \label{ssec:IntDis}
1601: 
1602: Instructive though the study of perfect interfaces may be in gaining
1603: an understanding of the role electronic structure mismatch may play
1604: in determining giant magnetoresistive effects, all measurements are
1605: made on devices which contain disorder, mostly in the diffusive regime.
1606: Because there is little information available from experiment about
1607: the nature of this disorder, it is very important to be able to model
1608: it in a flexible manner, introducing a minimum of free parameters. To
1609: model interfaces between materials with different lattice constants
1610: and disorder, we use the lateral supercells described in section
1611: \ref{ssec:Supercells}. Since this approach is formally only valid if
1612: sufficiently large supercells are used, we begin by studying how the
1613: interface conductance depends on the lateral supercell size.
1614: 
1615: \begin{figure}[tbp]
1616: \includegraphics[scale=0.35,clip=true]{cuco_size_2.669.eps}
1617: \caption{Interface conductance (in units of $10^{15} \:
1618: \Omega^{-1} \mathrm{m}^{-2}$) for a disordered Cu/Co (111)
1619: interface modelled as 2ML of 50\%-50\% alloy in a $\sqrt{H} \times
1620: \sqrt{H}$ lateral supercell as a function of 
1621: $\sqrt{H}$. The results are given for different randomly generated
1622: configurations of disorder (15 for minority spin, 5 for majority
1623: spin).
1624: Results are for a ``standard'' configuration.\cite{fn:standard} }
1625: \label{fig:config_av}
1626: \end{figure}
1627: 
1628: To perform fully self-consistent calculations for a number of large
1629: lateral supercells and for different configurations of disorder would
1630: be prohibitively expensive. Fortunately, the coherent potential
1631: approximation (CPA) is a very efficient way of calculating charge and
1632: spin densities for a substitutional disordered $A_x B_{1-x}$ alloy
1633: with an expense comparable to that required for an ordered system with
1634: a minimal unit cell.\cite{Soven:pr67} The output from such
1635: a calculation are atomic sphere potentials for the two sites,
1636: $\upsilon_A$ and $\upsilon_B$. The layer CPA approximation generalizes
1637: this to allow the concentration to vary from one layer to the next.
1638: \cite{Turek:97}
1639: 
1640: Once $\upsilon_A$ and $\upsilon_B$ have been calculated for some
1641: concentration $x$, an $H = H_1 \times H_2$ lateral supercell is
1642: constructed in which the potentials are distributed at random,
1643: maintaining the concentration for which they were self-consistently
1644: calculated. The conductances calculated for $4 \leq \sqrt{H} \leq 20 $ are shown in
1645: Fig.~\ref{fig:config_av} for a Cu/Cu(111) interface in which the Cu and
1646: the Co layers forming the interface are totally mixed to give two layers
1647: of 50\%-50\% interface alloy. For each value of $H$, the results for a
1648: number of different randomly generated disorder configurations are shown
1649: (20 for minority, 5 for minority spin). The sample to sample variation
1650: is largest for the minority spin case, ranging from $\pm 5 \%$ for a
1651: modest $4 \times 4$ unit cell and decreasing to less than $\pm 1 \%$
1652: for a $20 \times 20$ unit cell. For $\sqrt{H} \sim 10$, the spread in
1653: minority spin conductances is $\sim 5 \%$ which is comparable to the
1654: typical uncertainty we associated with the LDA error, the uncertainty
1655: in lattice constants or the error incurred by using the ASA.
1656: 
1657: Comparing now the conductances without and with disorder, we see that
1658: interface disorder has virtually no effect on the majority spin channel
1659: (0.434 versus $ 0.432 \times 10^{15} \: \Omega^{-1} \mathrm{m}^{-2}$)
1660: which is a consequence of the great similarity of the Cu and Co
1661: majority spin potentials and electronic structures. However, in the
1662: minority-spin channel the effect
1663: (0.364 versus $ 0.31 \times 10^{15} \: \Omega^{-1} \mathrm{m}^{-2}$) is
1664: much larger. As noted in the context of \eqref{eq:R_Schep}, a relatively
1665: small change in the interface transmission can lead to a large change in
1666: the interface resistance when account is taken of the finite conductance
1667: of the leads. We will return to the consequences for the spin-dependent
1668: interface resistance after completing the study of the interface
1669: transmission on which it is based.
1670: 
1671: When disorder is modelled in lateral supercells, the transmission
1672: probabilities can be classified as being {\em specular} or {\em diffuse}
1673: depending upon whether or not transverse momentum is
1674: conserved.\cite{Bruno:jmmm99,Drchal:prb02}
1675: In the presence of interface disorder, the conductance per unit area
1676: can be expressed as
1677: \begin{align}
1678: G &= G_{s} + G_{d}    \nonumber\\
1679:   &= \frac{e^2}{h} \sum_{\substack{ \mu \nu  \\
1680:                                   {\bf k}_{\parallel}  }}
1681: T_{\mu \nu } ({\bf k}_{\parallel}, {\bf k}_{\parallel})
1682:   + \frac{e^2}{h} \sum_{\substack{ \mu \nu  \\
1683:                                   {\bf k}_{\parallel} \neq {\bf k}_{\parallel}^{\prime }  }}
1684: T_{\mu \nu } ({\bf k}_{\parallel }, {\bf k}_{\parallel }^{\prime })
1685: \label{eq:sd}
1686: \end{align}%
1687: where ${\bf k}_{\parallel }$ and ${\bf k}_{\parallel }^{\prime }$ belong to the
1688: two dimensional Brillouin zone for ($1\times 1$) translational periodicity
1689: and
1690: $T_{\mu \nu }({\bf k_{\parallel }, k_{\parallel }^{\prime }})=
1691:  t_{\mu \nu }({\bf k_{\parallel }, k_{\parallel }^{\prime }})
1692:  t_{\mu \nu }^{\dagger }({\bf k_{\parallel }, k_{\parallel }^{\prime }})$.
1693: The transmission matrix elements between two Bloch states with the
1694: same ${\bf k_{\parallel }}$ are defined to be specular, those between
1695: scattering states with different ${\bf k_{\parallel }}$ as being diffuse.
1696: In the absence of interface disorder, there is by definition only a specular
1697: component.
1698: 
1699: \subsubsection*{Dependence of interface conductance on alloy concentration}
1700: 
1701: The results in Fig.~\ref{fig:config_av} were obtained for a structural
1702: model of the Co/Cu(111) interface consisting of two monolayers (2ML)
1703: of 50\%-50\% alloy that was derived from X-ray\cite{Henry:prb96}
1704: NMR\cite{deGronckel:prb91,Meny:prb92} and magnetic
1705: EXAFS\cite{Kapusta:jac99} studies.  Though the most plausible model
1706: there is at present, it contains large uncertainties. This makes it
1707: important to explore the consequences of varying the parameters
1708: defining the model. To do so, we calculate the conductance using 20x20
1709: lateral supercells as a function of alloy concentration for models in
1710: which the disorder is confined to one, two or four monolayers. The
1711: three models are defined in Fig.~\ref{fig:disord}. From the results
1712: shown in Fig.~\ref{fig:spec_diff}, it can be seen that the interface
1713: transmission for majority-spin electrons depends only very weakly on
1714: alloy concentration and its spatial distribution. The results for the
1715: 1ML, 2ML and 4ML models cannot be distinguished on the scale of the
1716: figure. When the conductance is decomposed using \eqref{eq:sd}, the
1717: diffuse component is found to be very small. Therefore, only the
1718: results for the minority-spin case need be examined in any detail.
1719: 
1720: \begin{figure}[tbp]
1721: \includegraphics[scale=0.7,clip=true]{int_dis.eps}
1722: \caption{Illustration of 3 different models of interface disorder
1723: considered.
1724: Top (1ML): disorder is modelled using one monolayer (ML) of [Cu$_{1-x}$Co${_x}$]
1725: alloy between Cu and Co leads, denoted as Cu$[$Cu$_{1-x}$Co$_{x}]$Co.
1726: Middle (2ML): disorder modelled in two MLs as
1727: Cu$[$Cu$_{1-x}$Co$_{x}|$Cu${_x}$Co$_{1-x}]$Co.
1728: Bottom (4ML): starting from the 2 ML disorder case, 1/3 of the
1729: concentration $x$ of impurity atoms is transferred to the next layer
1730: resulting in disorder in four MLs:
1731: Cu$[$Cu$_{1-\frac{x}{3}}$Co$_\frac{x}{3} | $Cu$_{1-\frac{2x}{3}}$Co$_{\frac{2x}{3}} |
1732: $Cu$_{\frac{2x}{3}}$Co$_{1-\frac{2x}{3}}|$Cu$_\frac{x}{3}$Co$_{1-\frac{x}{3}}]$Co.
1733: }
1734: \label{fig:disord}
1735: \end{figure}
1736: 
1737: We start by varying the alloy concentration over the full concentration
1738: range (0-100\%) for a disordered monolayer. The magnitude of resulting variation
1739: in transmission is limited ($\sim 7\%$) but exceeds the statistical
1740: spread between different configurations of disorder, which according
1741: to Fig.~\ref{fig:config_av} is less than $\pm 1 \%$.
1742: On adding Co to a layer of Cu, the transmission decreases, reaches a
1743: minimum for 10\% Co, then increases monotonically up to 70-90\% region
1744: where the transmission is \emph{higher} than for a clean
1745: interface.\cite{fn:xia01} 100\% Co represents a clean interface again,
1746: so this limit must yield the same transmission as 0\% Co.
1747: 
1748: The variation can be examined in terms of the specular and diffuse
1749: components defined in \eqref{eq:sd}. From Fig.~\ref{fig:spec_diff},
1750: it can be seen that, for the minority spin channel, the diffuse
1751: scattering by Co impurity atoms in Cu is stronger than that by Cu
1752: impurity atoms in Co. However, the specular scattering is also more
1753: strongly reduced by Cu in Co than by Co in Cu. The two effects largely
1754: cancel resulting in the observed undulatory total transmission as a
1755: function of the alloy concentration seen in the figure. The diffuse
1756: scattering has a maximum close to a 50\%-50\% alloy concentration where
1757: its contribution to the conductance is almost twice as large as from
1758: the specular scattering. While the conductance as such is scarcely
1759: effected, the strong diffuse scattering will play an important role in
1760: destroying the phase coherence of the electrons. Ultimately, this will
1761: be the physics underlying the so-called ``two current series resistor''
1762: (2CSR) model.\cite{Zhang:jap91,Lee:jmmm93,Valet:prb93}
1763: 
1764: \begin{figure}[t]
1765: \includegraphics[scale=0.35,clip=true]{G_tot_spec_diff_mz.eps}
1766: \caption{Interface conductance of a disordered Cu/Co (111)
1767: interface with disorder modelled in a $20 \times 20$ lateral
1768: supercell as a function of the alloy concentration $x$. Results
1769: are shown for the three different models described in
1770: Fig.~\ref{fig:disord} with disorder in 1, 2 or 4 MLs. Only a
1771: single disorder configuration was used and the size of the symbols
1772: corresponds to the spread in values found for this supercell size
1773: in Fig.~\ref{fig:config_av}. 
1774: For 1ML, the total conductance
1775: is resolved into specular and diffuse components (see the legend). 
1776: Results are for a ``standard'' configuration.\cite{fn:standard}}
1777: \label{fig:spec_diff}
1778: \end{figure}
1779: 
1780: If the disorder extends over more than a monolayer, then modelling the
1781: interface as several layers of homogeneous alloy is not obviously
1782: realistic. Instead, one might expect the layers closest to the interface
1783: to be most strongly mixed, the amount of mixing decreasing with the
1784: separation from the interface. A simple way to model this is to take
1785: two interface layers, one Cu and one Co, and to mix them in varying
1786: degrees. Denoting this Cu$|$Co interface as
1787: Cu$[$Cu$_{1-x}$Co$_{x}|$Cu${_x}$Co$_{1-x}]$Co
1788: we consider $0 \le x \le 0.5$ {\em i.e.}, the Cu concentration decreases
1789: monotonically from left to right. The calculated interface transmission
1790: is seen to essentially interpolate linearly the results obtained
1791: previously for the clean ($x=0$) and disordered ($x=0.5$) cases.
1792: 
1793: A slightly more elaborate model can be constructed from the 2ML model by
1794: distributing the $x$ impurity atoms so that $2x/3$ are in the interface
1795: layer while $x/3$ are to be found further from the original interface,
1796: in the following layer. This results in the concentration profile
1797: Cu$[$Cu$_{1-\frac{x}{3}}$Co$_\frac{x}{3} | $Cu$_{1-\frac{2x}{3}}$Co$_{\frac{2x}{3}} |
1798: $Cu$_{\frac{2x}{3}}$Co$_{1-\frac{2x}{3}}|$Cu$_\frac{x}{3}$Co$_{1-\frac{x}{3}}]$Co.
1799: $x=0$ corresponds to a completely ordered interface while the maximum
1800: value $x$ can have so that the concentration decreases from left to
1801: right monotonically is 75\%.
1802: This relatively small redistribution of intermixed atoms is seen to
1803: reduce the transmission by 15\% for $x=0.5$. Even for very good metals,
1804: relatively opaque interfaces can result when the electronic structure
1805: on either side have different characters. In such situations, disorder
1806: can influence the transmission strongly even reducing the interface
1807: resistance very substantially.\cite{Xia:prb01}
1808: A detailed analysis of the different contributions to the interface
1809: scattering in the 2ML and 4ML cases will be given in a separate
1810: publication.
1811: 
1812: \subsection{Analysis of Interface Disorder Scattering}
1813: \label{ssec:DefSca}
1814: 
1815: 
1816: \begin{figure}[b]
1817: \includegraphics[scale=1.0,angle=0,clip=true]{CuCo_111_maj_20x20_cm.eps}
1818: \caption{Fermi surface projections of majority-spin {\em fcc} Cu
1819: (a) and Co (b) derived from a single k-point using a $20 \times
1820: 20$ lateral supercell. The dark red point in the Cu Fermi surface
1821: projection corresponds to the point $Y$ in the top righthand panel
1822: of Fig.~\ref{fig:CuCo111_min}. $T(Y,\bf{k}_{\parallel}^{\prime })$
1823: is shown in (c), and in (d) magnified by a factor 500 where the
1824: ballistic component $T(Y,\bf{k}_{\parallel}^{\prime }=Y)$ is
1825: indicated by a white point because its value goes off the scale.
1826: Results are for a ``standard'' configuration\cite{fn:standard} 
1827: and averaged over 5 different configurations of disorder.}
1828: \label{fig:dis_scatt_maj}
1829: \end{figure}
1830: 
1831: The scattering induced by two layers of 50\%-50\% alloy is illustrated
1832: in Fig.~\ref{fig:dis_scatt_maj} and Fig.~\ref{fig:dis_scatt_min} for
1833: the majority and minority spins, respectively, of a Cu/Co(111)
1834: interface.  Calculations were performed for the single
1835: $\bf{k}_{\parallel}^{\mathbb{S}}$ point, $\Gamma $, and a $20 \times
1836: 20$ lateral supercell 
1837: equivalent to using a $1 \times 1$ interface
1838: cell and k-space sampling with $20 \times 20$ points in the
1839: corresponding BZ. Disorder averaging was carried out using 5 (for
1840: majority spin) or 20 (for minority spin) disorder configurations generated
1841: randomly.
1842: 
1843: Figs.~\ref{fig:dis_scatt_maj}(a) and \ref{fig:dis_scatt_maj}(b) show
1844: the majority-spin Fermi surface projections of {\em fcc} Cu and Co,
1845: respectively, obtained from ``unfolding'' the supercell calculation.
1846: The coarse $20 \times 20$ grid is seen to yield a good representation
1847: of the detailed Fermi surface projections shown in
1848: Fig.~\ref{fig:CuCo111_maj}.  $T(\bf{k}_{\parallel
1849: },\bf{k}_{\parallel}^{\prime })$ is shown in
1850: Fig.~\ref{fig:dis_scatt_maj}(c) for ${\bf k_{\parallel}} = Y$ on the
1851: $k_y$ axis in Fig.~\ref{fig:CuCo111_min}. Specular scattering
1852: dominates with $T( {\bf k}_{\parallel} = Y, {\bf
1853:   k}_{\parallel}^{\prime} = Y) = 0.93$.  The diffuse scattering is so
1854: weak that nothing can be seen on a scale of {\em T} from 0 to 1. To
1855: render it visible, a magnification by a factor 500 is needed,
1856: Fig.~\ref{fig:dis_scatt_maj}(d). The total diffuse scattering,
1857: $T_{d}(Y) = \sum_{ k_{\parallel }^{\prime} \neq k_{\parallel } }
1858: T({\bf k_{\parallel }}=Y, {\bf k_{\parallel }^{\prime}} \neq Y) =
1859: 0.04$ can be seen from the 
1860: figure to be made up of contributions of $T
1861: \sim 0.0004$ from roughly a quarter 
1862: of the BZ (100 ${\bf
1863:   k_{\parallel}}$ points) centred on ${\bf k_{\parallel }} = Y$. The
1864: total transmission, $T_{total}= T_{s} + T_{d} = 0.93 + 0.04 = 0.97$,
1865: compared to a transmission of 0.99 in the absence of disorder.  
1866: Similar results were obtained for other $\mathbf{k}_{||}$ points.
1867: In the majority case, there is thus a strong specular peak surrounded
1868: by a weak diffuse background.
1869: 
1870: 
1871: 
1872: 
1873: \begin{figure}[b]
1874: \includegraphics[scale=1.0,angle=0,clip=true]{CuCo_111_min_20x20_cm.eps}
1875: \caption{Fermi surface projections of minority-spin {\em fcc} Cu
1876: (a) and Co (b) derived from a single k-point using a $20 \times
1877: 20$ lateral supercell. The dark red point in the Cu Fermi surface
1878: projection corresponds to the point $Y^{\prime }$ in the top
1879: righthand panel of Fig.~\ref{fig:CuCo111_min}. (c)
1880: $T(Y,\mathbf{{k}_{\parallel }^{\prime })}$ and (d) $T(Y^{\prime
1881: },\mathbf{{k}_{\parallel }^{\prime })}$ calculated using 20
1882: different disorder configurations; the ballistic component
1883: $T(Y^{\prime },\mathbf{{k}_{\parallel }^{\prime }}=Y^{\prime })$
1884: is indicated by a white point because its value goes off scale.
1885: Results are for a ``standard'' configuration,\cite{fn:standard} 
1886: 20 different configurations of disorder were used.}
1887: \label{fig:dis_scatt_min}
1888: \end{figure}
1889: 
1890: The minority-spin Fermi surface projections of {\em fcc} Cu and Co are
1891: shown in Figs.~\ref{fig:dis_scatt_min}(a) and \ref{fig:dis_scatt_min}(b),
1892: respectively. Compared to the corresponding panels in
1893: Fig.~\ref{fig:CuCo111_min}, the $20 \times 20$ point representation is
1894: seen to be sufficient to resolve the individual Fermi surface sheets of
1895: Co. To study the effect of interface disorder, we consider scattering
1896: out of two different ${\bf k_{\parallel }}$s in Cu
1897: (Figs.~\ref{fig:dis_scatt_min}(c) and (d)). The first thing to
1898: note is the similarity of both transmission plots  to the projected FS
1899: of Co, Fig.~\ref{fig:dis_scatt_min}(b), suggesting very strong diffusive
1900: scattering proportional to the density of available final states.
1901: 
1902: The first case we consider is where ${\bf k_{\parallel}}=Y$ for
1903: which the transmission was zero as a result of the symmetry of the
1904: states along the $k_y$ axis in the absence of disorder.
1905: $T({\bf k_{\parallel }}= Y,{\bf k_{\parallel}^{\prime }})$ is shown in
1906: Fig.~\ref{fig:dis_scatt_maj}(c). By contrast with the majority-spin
1907: case just examined, there is now scattering to all other k-points in
1908: the 2D BZ,
1909: $\sum_{     k_{\parallel }^{\prime} \neq k_{\parallel }     }
1910: T({\bf k_{\parallel }}=Y, {\bf k_{\parallel }^{\prime}} \neq Y) = 0.58$
1911: while $T(Y,Y)$ has only increased from 0.00 in the clean case, to 0.01
1912: in the presence of disorder.
1913: The effect of disorder is to increase the total transmission,
1914: $T_{total}(Y)= \sum_{ k_{\parallel }^{\prime}  }
1915: T({\bf k_{\parallel }}=Y, {\bf k_{\parallel }^{\prime}})$ from
1916: 0.00 to $T_{s}(Y) + T_{d}(Y) = 0.01 + 0.58 = 0.59$; for states
1917: which were originally strongly reflected, disorder {\em increases} the
1918: transmission.
1919: 
1920: The second case we consider is that of a k-point slightly further away
1921: from the origin $\Lambda$ along the $k_y$ axis which had a high
1922: transmission, $T(Y^{\prime})= 0.98$, in the absence of disorder. For
1923: this k-point, $T({\bf k_{\parallel }}= Y^{\prime},{\bf
1924:   k_{\parallel}^{\prime }})$, shown in
1925: Fig.~\ref{fig:dis_scatt_maj}(d), looks very similar to
1926: Fig.~\ref{fig:dis_scatt_maj}(c). There is strong diffuse scattering
1927: with $\sum_{ k_{\parallel }^{\prime} \neq k_{\parallel } } T({\bf
1928:   k_{\parallel }}=Y^{\prime}, {\bf k_{\parallel }^{\prime}} \neq
1929: Y^{\prime}) = 0.54$ while $T(Y^{\prime},Y^{\prime})$ has been
1930: drastically decreased from 0.98 in the clean case, to 0.06 as a result
1931: of disorder. The total transmission, $T_{total}(Y^{\prime}) =
1932: T_{s}(Y^{\prime}) + T_{d}(Y^{\prime}) = 0.06 + 0.54 = 0.60$, is almost
1933: identical to what was found for the $Y$ point. The effect of disorder
1934: has been to {\em decrease} the transmission for states which were
1935: originally weakly reflected. The strong k-dependence of the
1936: transmission found in the specular case is largely destroyed by a
1937: small amount of disorder in the minority-spin channel. The
1938: contribution from specular component (integrated over 2D BZ) is reduced to 
1939: 15\% of the total transmission.
1940: 
1941: \subsection{Interface resistance}
1942: 
1943: To the best of our knowledge, spin-dependent interface transmissions
1944: have not yet been measured directly. What is usually done
1945: \cite{Pratt:prl91,Gijs:prl93} is to measure total resistances for a
1946: whole series of magnetic multilayers in which the total number of
1947: interfaces and/or the thicknesses of the individual layers is varied.
1948: The measured results are interpreted in terms of volume resistivities
1949: and interface resistances. By applying an external magnetic field, the
1950: magnetizations of neighbouring layers which are oriented antiparallel
1951: (AP) can be forced to line up in parallel (P). By measuring the
1952: resistances in both cases, spin-dependent volume resistivities and
1953: interface resistances can be extracted using the two current series
1954: resistor model.\cite{Zhang:jap91,Lee:jmmm93,Valet:prb93} If we take
1955: expression \eqref{eq:R_Schep} which relates the interface transmission
1956: to the interface resistance occurring in the 2CSR model as given,%
1957: \cite{Schep:prb97,Stiles:prb00} we can study how typical uncertainties in interface
1958: transmission, arising from arbitrary assumptions about the interface
1959: disorder, lattice constant or basis set translate into uncertainty in
1960: predicted interface resistances.
1961: Using the transmission probabilities from Fig.~\ref{fig:spec_diff} in
1962: \eqref{eq:R_Schep} results in the curves shown in Fig.~\ref{fig:int_res}.
1963: For comparison, a range of literature values for the spin-dependent
1964: interface resistances derived from experiments on sputtered and MBE
1965: (molecular beam exitaxy) grown multilayers\cite{Bass:jmmm99} is
1966: included in the figure.
1967: 
1968: \begin{table}[b]
1969: \begin{ruledtabular}
1970: \begin{tabular}{lccc}
1971: \multicolumn{1}{c}{$a (\mathring{A}$)} &
1972:        \multicolumn{2}{c} {3.549} &
1973:                                      \multicolumn{1}{c} {3.614} \\
1974: \hline
1975: \multicolumn{1}{c}{Basis} &
1976:     \multicolumn{1}{c}{\em spdf} &
1977:                       \multicolumn{1}{c}{\em spd} &
1978:                                                  \multicolumn{1}{c}{\em spd} \\
1979: \hline
1980: $R^{\rm maj}(111)$  &  0.46 &  0.39 &  0.34  \\
1981: $R^{\rm min}(111)$  &  1.33 &  1.32 &  1.37
1982: \end{tabular}
1983: \end{ruledtabular}
1984: \caption{Interface resistances, in units of f$ \Omega \mathrm{m}^{2}$,
1985: for ordered interfaces, calculated using expression \eqref{eq:R_Schep}
1986: and the data from Tables \ref{tab:A} and \ref{tab:B}.
1987: The values given here for a lattice constant of $a=3.614\mathring{A}$ differ
1988: slightly from those reported in Ref.~\onlinecite{Xia:prb01} which were
1989: performed using energy-independent muffin-tin orbitals linearized about
1990: the centers of gravity of the occupied conduction states and not at the
1991: Fermi energy. The current implementation\cite{Zwierzycki:prb03} uses
1992: energy-dependent, (non-linearized) MTO's, calculated exactly at the Fermi
1993: energy which improves the accuracy at no additional cost.
1994: }
1995: \label{tab:C}
1996: \end{table}
1997: 
1998: For the minority-spin case, experimental values (in units of
1999: f$ \Omega \mathrm{m}^{2}$) range from 1.30-1.80 compared to calculated
2000: values of
2001: 1.29 for Cu$[$Cu$_{.3}$Co$_{.7}]$Co,
2002: through 1.37 for a disorder-free interface,
2003: to a value of 2.25 for the 4ML model with $x=0.5$,
2004: Cu$[$Cu$_{.83}$Co$_{.17}|$Cu$_{.67}$Co$_{.33}|$Cu$_{.33}$Co$_{.67}|$Cu$_{.17}$Co$_{.83}]$Co.
2005: The influence of lattice constant and basis set on the clean interface
2006: resistance values is small (see Table~\ref{tab:C}).
2007: The present modelling of interface alloying shows that the interface
2008: resistance is more strongly dependent on the detailed spatial
2009: distribution of disorder than was previously found\cite{Xia:prb01}
2010: where only the concentration range $x=0.5 \pm 0.06$ of the 2ML interface
2011: alloy model extracted from experiment%
2012: \cite{deGronckel:prb91,Meny:prb92,Henry:prb96,Kapusta:jac99}
2013: was explored.
2014: 
2015: \begin{figure}[t]
2016: \includegraphics[scale=0.35,clip=true]{res_mz.eps}
2017: \caption{Interface resistance for disordered interfaces as a
2018: function of the alloy concentration used to model disordered
2019: interfaces calculated using \eqref{eq:R_Schep} and the
2020: transmission probabilities shown in Fig.~\ref{fig:spec_diff}. The
2021: experimental values for sputtered and MBE grown multilayers cited
2022: in Table I of Ref.~\onlinecite{Bass:jmmm99} span a range of values
2023: which is indicated by the shaded regions.}
2024: \label{fig:int_res}
2025: \end{figure}
2026: 
2027: For the majority-spin case, the spread in values of the interface
2028: resistance extracted from experiment (for the same samples as for
2029: the minority-spin case) is quite small, 0.22-0.25, and does not
2030: overlap with the values of $0.34$
2031: found for a lattice constant of $a=3.614\mathring{A}$. Unlike the
2032: minority-spin case, changing the lattice constant or using an $spdf$
2033: basis leads to substantially {\em larger} values (Table~\ref{tab:C}).
2034: Because the majority-spin transmission does not depend on the details
2035: of the interface disorder, this cannot be the origin of the discrepancy.
2036: Motivated by the weak scattering in this case, we examine the validity
2037: \cite{Valet:prb93,Gijs:ap97,Bass:jmmm99,Tsymbal:prb00b,Shpiro:prb00}
2038: of the 2CSR model by calculating the resistance of a magnetic multilayer
2039: containing  a large number of disordered interfaces and plot the
2040: resistance added by each additional interface in Fig.~\ref{fig:diff_res}.
2041: Compared to similar calculations in Ref.~\onlinecite{Xia:prb01},
2042: the number of interfaces, size of lateral supercell ($10 \times 10$)
2043: and disorder configurations averaged over are increased substantially.
2044: While the calculations are in very good agreement with Ohm's law for
2045: the strongly scattering minority-spin case, it can be seen that this
2046: is not the case for the majority-spin electrons. For a small number of
2047: interfaces there is a clear breakdown of Ohm's law and thus of the 2CSR
2048: model. The interface resistance eventually saturates at a value much
2049: lower than those extracted from experiment. While inclusion of bulk
2050: scattering will modify this picture somewhat, exploratory calculations%
2051: \cite{Gerritsen:02} indicate that the type of ``bulk'' impurities which
2052: may be reasonably expected to be found in sputtered or MBE grown
2053: multilayers affect the minority spin electrons much more than the
2054: majority spins. Agreement for the latter can only be achieved at the
2055: expense of ruining good agreement for the former.
2056: 
2057: \begin{figure}[btp]
2058: \includegraphics[scale=0.35,clip=true]{diff_intf_mz.eps}
2059: \caption{Differential interface resistance as the number of
2060: interfaces increase for a disordered Cu/Co(111) multilayer
2061: embedded between Cu leads. A $10 \times 10$ lateral supercell was
2062: used and the interface was modelled as two layers of 50\%-50\%
2063: alloy (2ML model). The results represent an average over 5 disorder configurations
2064: and were obtained for a ``standard'' configuration.\cite{fn:standard}
2065: The range of experimental values\cite{Bass:jmmm99} is indicated by the shaded regions.
2066: }
2067: \label{fig:diff_res}
2068: \end{figure}
2069: 
2070: \section{DISCUSSION}
2071: \label{sec:discussion}
2072: 
2073: Details of a muffin tin orbital-based method suitable for calculating
2074: from first-principles scattering matrices involving layered magnetic
2075: materials have been given. In a wide range of applications,%
2076: \cite{Xia:prb01,Xia:prb02,Xia:prl02,Zwierzycki:prb03,Bauer:prl04,Zwierzycki:prb05}
2077: it has been shown to be much more efficient and transparent than a
2078: previously used LAPW-based method.%
2079: \cite{vanHoof:97,Schep:prb97,vanHoof:prb99}
2080: Various other schemes have been developed for calculating the
2081: transmission of electrons through an interface (or a more extended
2082: scattering region)
2083: both from first principles,\cite{Stiles:prb88,Stiles:prl91,vanHoof:97,%
2084: MacLaren:prb99,Kudrnovsky:prb00,Riedel:prb01,Taylor:prb01,%
2085: Brandbyge:prb02,Wortmann:prb02a,Wortmann:prb02b,Thygesen:prb03,%
2086: Mavropoulos:prb04,Khomyakov:prb04} or using as input electronic
2087: structures which were calculated from first principles.%
2088: \cite{Tsymbal:jp97,Mathon:prb97a,Mathon:prb97c,Sanvito:prb99,%
2089: Butler:prb01,Mathon:prb01,Mathon:prb05}
2090: Most are based upon a formulation for the conductance in terms of
2091: non-equilibrium Green's functions \cite{Caroli:jpc71a} (NEGF) which
2092: reduces in the appropriate limit to the well known Fisher-Lee (FL)
2093: linear-response form\cite{Fisher:prb81} for the conductance of a finite
2094: disordered wire embedded between crystalline leads. Most implementations
2095: of the NEGF or FL schemes have two disadvantages. (i) The transmission
2096: is calculated for a complex energy which leads to difficulties in
2097: studying for example, tunneling magnetoresistance, where the finite
2098: imaginary part can give rise to an exponential decay which obscures
2099: the interesting physical decay of the transmission as a function of
2100: the barrier thickness. (ii) For a given value of transverse crystal
2101: momentum, the transmission is expressed as a trace over the basis set
2102: in terms of which the Green's function and self-energy are expressed.%
2103: \cite{Khomyakov:prb05} While this has the advantage that the total
2104: transmission can be calculated without explicitly determining the
2105: scattering states and can be computationally efficient, summation of
2106: the contributions from multiple scattering states can obscure real
2107: physical effects, for example, the role of the symmetries of individual
2108: scattering states seen in Fig.~\ref{fig:CuCo111_min}. Explicit
2109: determination of the scattering states not only makes a detailed
2110: analysis of the scattering possible. The full scattering matrix,
2111: expressed in terms of the scattering states, can be used to bridge%
2112: \cite{Xia:prl02} the gap between first-principles electronic structure
2113: calculations and phenomenological models of transport used to analyse
2114: complex situations where a full first-principles treatment is not
2115: practical.
2116: 
2117: We have instead made use of an alternative technique, suitable for
2118: Hamiltonians that can be represented in tight-binding form, that was
2119: formulated by Ando\cite{Ando:prb91} and is based upon direct matching
2120: of the scattering-region wave function to the Bloch modes of the leads.
2121: The relationship between the wave function matching \cite{Ando:prb91}
2122: and Green function \cite{Caroli:jpc71a,Fisher:prb81} approaches is not
2123: immediately obvious. It was suggested recently that WFM was incomplete
2124: \cite{Krstic:prb02} but the equivalence of the two approaches could be
2125: proven.\cite{Khomyakov:prb05} Schemes similar in spirit to our own, but
2126: based upon empirical tight-binding Hamiltonians have been presented by
2127: Sanvito {\em et al.}\cite{Sanvito:prb99} and by Velev.%
2128: \cite{Velev:prb03,Velev:prb04} In contrast to these schemes, our TB-MTO
2129: formalism is a parameter-free approach that has all of the advantages
2130: derived from self-consistent determination of potentials and spin
2131: densities for systems for which these are not known from experiment.
2132: Judging from the size of systems to which it has been applied, it would
2133: seem that our implementation is nevertheless substantially more efficient
2134: than these empirical schemes. The scattering regions treated in
2135: Figs.~\ref{fig:config_av},\ref{fig:spec_diff},\ref{fig:dis_scatt_maj} and
2136: \ref{fig:dis_scatt_min} contained as many as 3200 atoms
2137: ($20 \times 20$ lateral supercell $\times $ 8 principal layers where the
2138: potential was allowed to deviate from its bulk values) or, in the case of
2139: Fig.~\ref{fig:diff_res},
2140: $\sim 15000$ atoms ($10 \times 10$ lateral supercell $\times$ 150
2141: principal layers).
2142: Our WFM scheme should not be confused\cite{Velev:prb04} with a recently
2143: developed transport formalim\cite{Kudrnovsky:prb00,Drchal:prb02} also
2144: based upon TB-LMTOs but which makes use of the Caroli NEFG expression
2145: for the conductance in terms of a trace and a complex energy. Khomyakov
2146: and Brocks\cite{Khomyakov:prb04} have developed a scheme analogous to
2147: ours but based upon pseudopotentials and a real space grid which make
2148: it more suitable for studying quantum wires or the type of open
2149: structures studied in molecular electronics, but is computationally
2150: much more expensive.
2151: 
2152: A third approach based upon
2153: ``embedding''\cite{Inglesfield:jpc81,Crampin:jp92} has been combined
2154: with full-potential linearized augmented plane wave method to
2155: yield what is probably the most accurate scheme to date
2156: \cite{vanHoof:97,Wortmann:prb02a,Wortmann:prb02b} but like the real
2157: space grid WFM method,\cite{Khomyakov:prb04} these methods are
2158: numerically very demanding.
2159: 
2160: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XSUMMARY
2161: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XSUMMARY
2162: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XSUMMARY
2163: 
2164: \section{SUMMARY}
2165: \label{sec:summary}
2166: 
2167: Details of a wave-function matching method suitable for calculating the
2168: scattering matrices in magnetic metallic hybrid structures based upon
2169: first-principles tight-binding muffin tin orbitals have been given and
2170: illustrated with calculations for a variety of Co/Cu(111) interface-related
2171: problems. The minimal basis of localized orbitals is very efficient, allowing
2172: large lateral supercells to be handled. This allow us to model materials with
2173: large lattice mismatch or to study transport in the diffusive regime. Because
2174: the scattering states are calculated explicitly, the effect of various types
2175: of scattering can be analyzed in detail.
2176: 
2177: \begin{acknowledgments}
2178: This work is part of the research program for the
2179: ``Stichting voor Fundamenteel Onderzoek der Materie'' (FOM)
2180: and the use of supercomputer facilities was sponsored by the
2181: ``Stichting Nationale Computer Faciliteiten'' (NCF), both financially
2182: supported by the
2183: ``Nederlandse Organisatie voor Wetenschappelijk Onderzoek'' (NWO).
2184: It was also supported by the European Commission's
2185: Research Training Network {\em Computational Magnetoelectronics}
2186: (contract No. HPRN-CT-2000-00143) as well as by the
2187: NEDO International Joint Research program {\em Nano-scale Magnetoelectronics}.
2188: MZ wishes also to acknowledge support from KBN grant
2189: No.~PBZ-KBN-044/P03-2001.
2190: We are grateful to: Ilya Turek for his TB-LMTO-SGF layer code which we
2191: used to generate self-consistent potentials and for numerous discussions
2192: about the method; Anton Starikov for permission to use his version of the
2193: TB-MTO code based upon sparse matrix techniques to perform some of the
2194: calculations.
2195: \end{acknowledgments}
2196: 
2197: 
2198: \appendix
2199: 
2200: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XVELOCITIES
2201: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XVELOCITIES
2202: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XVELOCITIES
2203: \section{Velocities}
2204: \label{sec:velocities}
2205: 
2206: Expressions for the velocities of the propagating modes in the
2207: leads (Sect.~\ref{ssec:Leads}) are more easily derived using an
2208: energy independent Hamiltonian than the energy dependent
2209: tail-cancellation condition of section \ref{ssec:MTOs}. To do so,
2210: we make use of the close relationship between the KKR
2211: tail-cancellation equation (\ref{eq:TBPmS}) and the linearized MTO
2212: (LMTO) Hamiltonian, both of which can be expressed in terms of the
2213: Hermitian matrix\cite{Andersen:85,Andersen:prb86}
2214: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
2215: \begin{gather}
2216:   h^\alpha(\varepsilon)
2217:   =- [\dot{P}^\alpha(\varepsilon)]^{-1/2}
2218:       \left(P^\alpha(\varepsilon)-S^\alpha \right)
2219:      [\dot{P}^\alpha(\varepsilon)]^{-1/2}          \nonumber \\
2220:   = -P^\alpha(\varepsilon) [\dot{P}^\alpha(\varepsilon)]^{-1}
2221:   + [\dot{P}^\alpha(\varepsilon)]^{-1/2} S^\alpha
2222:                [\dot{P}^\alpha(\varepsilon)]^{-1/2}.
2223:   \label{eq:halpha}
2224: \end{gather}
2225: Fixing the energy at $\varepsilon=\varepsilon_F$
2226: and defining the potential parameters\cite{Andersen:prb86,Andersen:87}
2227: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
2228: \begin{subequations}
2229: \label{eq:potpar}
2230: \begin{equation}
2231: \sqrt{d^\alpha}=[\dot{P}^\alpha(\varepsilon_F)]^{-1/2}
2232: \label{subeq:1}
2233: \end{equation}
2234: \begin{equation}
2235: c^\alpha=-P^\alpha(\varepsilon_F)/ \dot{P}^\alpha(\varepsilon_F) + \varepsilon_F,
2236: \end{equation}
2237: \end{subequations}
2238: %
2239: (\ref{eq:halpha}) can be written as
2240: \begin{equation}
2241:  h^\alpha \equiv h^\alpha(\varepsilon_F) = c^\alpha
2242:           + \sqrt{d^\alpha} \; S^\alpha \, \sqrt{d^\alpha}-\varepsilon_F
2243:   \label{eq:tbham}
2244: \end{equation}
2245: %
2246: Equation (\ref{eq:tbham}) has the form of a two-center tight binding
2247: Hamiltonian whose energy is given relative to $\varepsilon_F$. It
2248: provides the lowest order approximation\cite{Andersen:85,Andersen:prb86}
2249: to the full LMTO Hamiltonian and yields eigenvalues correct to first order
2250: in $(\varepsilon-\varepsilon_F)$.
2251: For eigenvalues equal to $\varepsilon_F$, it yields eigenvectors which
2252: are equal to those determined by the tail-cancellation condition
2253: (\ref{eq:TBPmS}), up to a scaling factor $(\dot{P}^\alpha)^{-1/2}$.
2254: 
2255: To calculate the group velocities of states precisely at the
2256: linearization energy, in the present case at the Fermi energy,
2257: the first-order Hamiltonian (\ref{eq:tbham}) can be used since any
2258: error vanishes identically for $\varepsilon(\mathbf{k})=\varepsilon_F$.
2259: Using the translational symmetry of the leads, the Hamiltonian
2260: (\ref{eq:tbham}) for Bloch vector ${\bf k}$ is
2261: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
2262: \begin{equation}
2263:   h^\alpha_{RL,R'L'}(\mathbf{k})=
2264:   \sum_{\mathbf{T}}e^{i\mathbf{k}\cdot\mathbf{T}}
2265:   h^\alpha_{RL,(R'+T)L'}
2266:   \label{eq:tbham_bloch}
2267: \end{equation}
2268: where $RL$ labels the sites and orbitals within the unit cell
2269: and $\mathbf{T}$ runs over lattice vectors.
2270: The energy eigenvalues $\varepsilon_{\mu}(\mathbf{k})$ are the
2271: expectation values
2272: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
2273: \begin{equation}
2274:   \varepsilon_{\mu}(\mathbf{k})=\mathbf{a}^\dagger_{\mu}(\mathbf{k})
2275:   h^{\alpha}(\mathbf{k}) \mathbf{a}_{\mu}(\mathbf{k})
2276:   \label{eq:expect_val}
2277: \end{equation}
2278: where the eigenvectors $\mathbf{a}_{\mu}(\mathbf{k})$ are indexed by $RL$ and
2279: we assumed normalization $\mathbf{a}^\dagger_{\mu}\cdot\mathbf{a}_{\mu}=1$.
2280: It is now straightforward to calculate the group velocity of the propagating
2281: mode
2282: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%=====================
2283: \begin{equation}
2284:   \begin{split}
2285:  {{\bm \upsilon }}_{\mu}=\frac{1}{\hbar}
2286:   \frac{\partial \varepsilon_{\mu}(\mathbf{k})}{\partial \mathbf{k}}
2287:   & = \frac{i}{\hbar}\sum_{\mathbf{T}}
2288:                     \mathbf{T}e^{i \mathbf{k} \cdot \mathbf{T}} \times  \\
2289:   & \sum_{RL,R'L'}  a^{*}_{RL}  h^{\alpha}_{RL,(R'+T)L'} a_{R'L'} \\
2290:   \end{split}
2291:   \label{eq:3dvelocity}
2292: \end{equation}
2293: In the mixed representation $\left|I,\mathbf{k}_{\parallel}\right>$ defined
2294: in section \ref{ssec:Leads} \eqref{eq:3dvelocity} gives for the
2295: velocity in the stacking direction
2296: \begin{equation}
2297:   \upsilon_{\mu}=\frac{id}{\hbar}\left[
2298:     \mathbf{a}^\dagger_{\mu} h^\alpha_{I,I+1}(\mathbf{k}_{\parallel})
2299:     \lambda_{\mu} \mathbf{a}_{\mu}-\mathbf{h.c.}
2300:     \right]
2301:   \label{eq:layered_velociy}
2302: \end{equation}
2303: where $d$ is the distance between equivalent monolayers in adjacent
2304: principal layers (PL), the hopping is assumed (as in section
2305: \ref{ssec:Leads}) to extend only between neighbouring PLs and
2306: $\lambda_{\mu}=exp(i\mathbf{k}\cdot\mathbf{T}^0)$ with $\mathbf{T}^0$
2307: connecting equivalent sites in the neighbouring PLs.
2308: Using the definition \eqref{eq:halpha} of $h^\alpha$  and
2309: recalling that the solutions $\mathbf{u}_{\mu}$ of the
2310: tail-cancellation equation \eqref{eq:TBPmS} take implicitly into
2311: account the scaling factor $(\dot{P}^\alpha)^{-1/2}$ we arrive at
2312: equation \eqref{eq:velocity}.
2313: 
2314: 
2315: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XSYMMETRY RELATIONS
2316: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XSYMMETRY RELATIONS
2317: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% XSYMMETRY RELATIONS
2318: \section{Symmetry relations}
2319: \label{sec:symmetry}
2320: 
2321: If we look closely at the transmission probabilities in
2322: Fig.~\ref{fig:CuCo111_min}, we see that the sheet resolved
2323: transmissions exhibit the geometrical symmetry of the underlying
2324: lattice (\emph{i.e.} the three-fold rotational axis). The total
2325: transmission probability on the other hand possesses an extra
2326: inversion symmetry,
2327: $T({\bf k_{\parallel }})=T(-{\bf k_{\parallel }})$, which results in
2328: plots with a six-fold rotational axis. This higher symmetry is the
2329: manifestation of the fundamental time-reversal symmetry obeyed in
2330: the absence of spin-orbit coupling and a magnetic field. In
2331: the case of the bulk system time-reversal symmetry grants that for
2332: every eigenstate $\psi_\alpha({\bf k})$ there exists the
2333: counterpart with the same energy and opposite wave vector
2334: (\emph{i.e.} $\varepsilon_\alpha({\bf k})=\varepsilon_\alpha(-{\bf
2335: k})$) and the wave functions are related by the complex conjugate.
2336: The situation is more complicated in the case of the scattering
2337: state. Consider a state incoming from the
2338: left lead and  scattered in the middle region. The wave function
2339: consists then of the incoming and reflected states in the left
2340: lead
2341: \begin{equation}
2342:   \Psi^r_{\mathcal{L}}({\bf k_{\parallel }})=\psi^{+}_{\mu}({\bf k_{\parallel }})+
2343:   \sum_{\mu'}r_{\mu'\mu}({\bf k_{\parallel }})\psi^{-}_{\mu'}({\bf k_{\parallel }})
2344:   \label{eq:left_ret}
2345: \end{equation}
2346: and of the transmitted states in the right lead
2347: \begin{equation}
2348:   \Psi^{r}_{\mathcal{R}}({\bf k_{\parallel }})
2349:   =\sum_{\nu}t_{\nu\mu}({\bf k_{\parallel }})\psi^+_{\nu}({\bf k_{\parallel }}).
2350:   \label{eq:right_ret}
2351: \end{equation}
2352: The time reversal operation transforms the above ``retarded'' state into the
2353: ``advanced'' one in which a number of  incoming states (from the left
2354: and the right) combine to produce a single outgoing state on the
2355: left, \emph{i.e.}
2356: \begin{equation}
2357:   \Psi^a_{\mathcal{L}}(-{\bf k_{\parallel }})=
2358:   \sum_{\mu'}r^{*}_{\mu'\mu}({\bf k_{\parallel }})\psi^{+}_{\mu'}(-{\bf k_{\parallel }})
2359:   +\psi^{-}_{\mu}(-{\bf k_{\parallel }})
2360:   \label{eq:left_adv}
2361: \end{equation}
2362: and
2363: \begin{equation}
2364:   \Psi^{a}_{\mathcal{R}}(-{\bf k_{\parallel }})=
2365:   \sum_{\nu}t^{*}_{\nu\mu}({\bf k_{\parallel }})\psi^-_{\nu}(-{\bf k_{\parallel }}).
2366:   \label{eq:right_adv}
2367: \end{equation}
2368: Equations (\ref{eq:left_adv}) and (\ref{eq:right_adv}) impose a set
2369: of conditions on the values of scattering coefficients for the states
2370: with $-{\bf k_{\parallel }}$. Combined with the analogous conditions derived
2371: for the states with the incoming state in the right lead, they are
2372: compactly expressed as
2373: \begin{equation}
2374:   I=S\left(-{\bf k_{\parallel }}\right)S^{*}\left({\bf k_{\parallel }}\right)\;\;\Rightarrow\;\;
2375:   S\left(-{\bf k_{\parallel }}\right)=S^{T}\left({\bf k_{\parallel }}\right).
2376: \label{eq:scatt_symm}
2377: \end{equation}
2378: The scattering matrix $S$ is defined as
2379: \begin{equation}
2380:   S=\left(
2381:     \begin{array}{cc}
2382:       r & t' \\
2383:       t & r'
2384:     \end{array}
2385:   \right)
2386:   \label{eq:scatt_def}
2387: \end{equation}
2388: where $r^{(}{'}^{)}$ and  $t^({'}^)$ are matrices in the space of the lead
2389: modes and the primed coefficients describe scattering of the states
2390: incoming from the right. More specifically we have:
2391: \begin{equation}
2392:   t_{\nu\mu}(-{\bf k_{\parallel }})=t'_{\mu\nu}({\bf k_{\parallel }})
2393:   \;\;\mathrm{and}\;\;
2394:   r_{\mu'\mu}(-{\bf k_{\parallel }})=r_{\mu\mu'}({\bf k_{\parallel }})
2395:   \label{eq:tr_symm}
2396: \end{equation}
2397: Equation~\eqref{eq:tr_symm} gives
2398: \begin{equation}
2399:   T_{\mathcal{LR}}(-{\bf k_{\parallel }})
2400:  =\sum_{\nu\mu} |t_{\nu\mu}(-{\bf k_{\parallel }})|^2
2401:  =\sum_{\mu\nu}|t'_{\mu\nu}({\bf k_{\parallel }})|^2
2402:  =T_{\mathcal{RL}}({\bf k_{\parallel }})
2403: \end{equation}
2404: In addition, for any two-terminal device, the Hermiticity of the
2405: scattering matrix guarantees that
2406: $T_{\mathcal{RL}}({\bf k_{\parallel }})=T_{\mathcal{LR}}({\bf k_{\parallel}})$
2407: (see Ref.~\onlinecite{Datta:95}) which finally proves the in-plane
2408: inversion symmetry mentioned at the beginning. The last step can
2409: not however be taken for the partial (FS resolved) transmission
2410: probabilities. These quantities thus possess only the geometrical
2411: symmetry of the system.
2412: 
2413: \begin{thebibliography}{85}
2414: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
2415: \expandafter\ifx\csname bibnamefont\endcsname\relax
2416:   \def\bibnamefont#1{#1}\fi
2417: \expandafter\ifx\csname bibfnamefont\endcsname\relax
2418:   \def\bibfnamefont#1{#1}\fi
2419: \expandafter\ifx\csname citenamefont\endcsname\relax
2420:   \def\citenamefont#1{#1}\fi
2421: \expandafter\ifx\csname url\endcsname\relax
2422:   \def\url#1{\texttt{#1}}\fi
2423: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
2424: \providecommand{\bibinfo}[2]{#2}
2425: \providecommand{\eprint}[2][]{\url{#2}}
2426: 
2427: \bibitem[{\citenamefont{Schep et~al.}(1995)\citenamefont{Schep, Kelly, and
2428:   Bauer}}]{Schep:prl95}
2429: \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Schep}},
2430:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}}, \bibnamefont{and}
2431:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2432:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{74}},
2433:   \bibinfo{pages}{586} (\bibinfo{year}{1995}).
2434: 
2435: \bibitem[{\citenamefont{Zahn et~al.}(1995)\citenamefont{Zahn, Mertig, Richter,
2436:   and Eschrig}}]{Zahn:prl95}
2437: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Zahn}},
2438:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Mertig}},
2439:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Richter}}, \bibnamefont{and}
2440:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Eschrig}},
2441:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{75}},
2442:   \bibinfo{pages}{2996} (\bibinfo{year}{1995}).
2443: 
2444: \bibitem[{\citenamefont{Weinberger et~al.}(1996)\citenamefont{Weinberger, Levy,
2445:   Banhare, Szunyogh, and Ujfalussy}}]{Weinberger:jp96}
2446: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Weinberger}},
2447:   \bibinfo{author}{\bibfnamefont{P.~M.} \bibnamefont{Levy}},
2448:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Banhare}},
2449:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Szunyogh}}, \bibnamefont{and}
2450:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Ujfalussy}},
2451:   \bibinfo{journal}{J. Phys.: Condens. Matter.} \textbf{\bibinfo{volume}{8}},
2452:   \bibinfo{pages}{7677} (\bibinfo{year}{1996}).
2453: 
2454: \bibitem[{\citenamefont{Schep et~al.}(1997)\citenamefont{Schep, van Hoof,
2455:   Kelly, Bauer, and Inglesfield}}]{Schep:prb97}
2456: \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Schep}},
2457:   \bibinfo{author}{\bibfnamefont{J.~B. A.~N.} \bibnamefont{van Hoof}},
2458:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2459:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2460:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~E.}
2461:   \bibnamefont{Inglesfield}}, \bibinfo{journal}{Phys. Rev. B}
2462:   \textbf{\bibinfo{volume}{56}}, \bibinfo{pages}{10805} (\bibinfo{year}{1997}).
2463: 
2464: \bibitem[{\citenamefont{Zahn et~al.}(1998)\citenamefont{Zahn, Binder, Mertig,
2465:   Zeller, and Dederichs}}]{Zahn:prl98}
2466: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Zahn}},
2467:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Binder}},
2468:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Mertig}},
2469:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Zeller}}, \bibnamefont{and}
2470:   \bibinfo{author}{\bibfnamefont{P.~H.} \bibnamefont{Dederichs}},
2471:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{80}},
2472:   \bibinfo{pages}{4309} (\bibinfo{year}{1998}).
2473: 
2474: \bibitem[{\citenamefont{van Hoof et~al.}(1999)\citenamefont{van Hoof, Schep,
2475:   Brataas, Bauer, and Kelly}}]{vanHoof:prb99}
2476: \bibinfo{author}{\bibfnamefont{J.~B. A.~N.} \bibnamefont{van Hoof}},
2477:   \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Schep}},
2478:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Brataas}},
2479:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2480:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2481:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{59}},
2482:   \bibinfo{pages}{138} (\bibinfo{year}{1999}).
2483: 
2484: \bibitem[{\citenamefont{MacLaren et~al.}(1999)\citenamefont{MacLaren, Zhang,
2485:   Butler, and Wang}}]{MacLaren:prb99}
2486: \bibinfo{author}{\bibfnamefont{J.~M.} \bibnamefont{MacLaren}},
2487:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Zhang}},
2488:   \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Butler}},
2489:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Wang}},
2490:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{59}},
2491:   \bibinfo{pages}{5470} (\bibinfo{year}{1999}).
2492: 
2493: \bibitem[{\citenamefont{Kudrnovsk\'{y}
2494:   et~al.}(2000)\citenamefont{Kudrnovsk\'{y}, Drchal, Blaas, Weinberger, Turek,
2495:   and Bruno}}]{Kudrnovsky:prb00}
2496: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kudrnovsk\'{y}}},
2497:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Drchal}},
2498:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Blaas}},
2499:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Weinberger}},
2500:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Turek}}, \bibnamefont{and}
2501:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Bruno}},
2502:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{62}},
2503:   \bibinfo{pages}{15084} (\bibinfo{year}{2000}).
2504: 
2505: \bibitem[{\citenamefont{Xia et~al.}(2001)\citenamefont{Xia, Kelly, Bauer,
2506:   Turek, Kudrnovsk\'{y}, and Drchal}}]{Xia:prb01}
2507: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Xia}},
2508:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2509:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2510:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Turek}},
2511:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kudrnovsk\'{y}}},
2512:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Drchal}},
2513:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{63}},
2514:   \bibinfo{pages}{064407} (\bibinfo{year}{2001}).
2515: 
2516: \bibitem[{\citenamefont{Riedel et~al.}(2001)\citenamefont{Riedel, Zahn, and
2517:   Mertig}}]{Riedel:prb01}
2518: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Riedel}},
2519:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Zahn}}, \bibnamefont{and}
2520:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Mertig}},
2521:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{63}},
2522:   \bibinfo{pages}{195403} (\bibinfo{year}{2001}).
2523: 
2524: \bibitem[{\citenamefont{Taylor et~al.}(2001)\citenamefont{Taylor, Guo, and
2525:   Wang}}]{Taylor:prb01}
2526: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Taylor}},
2527:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Guo}}, \bibnamefont{and}
2528:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Wang}},
2529:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{63}},
2530:   \bibinfo{pages}{245407} (\bibinfo{year}{2001}).
2531: 
2532: \bibitem[{\citenamefont{Brandbyge et~al.}(2002)\citenamefont{Brandbyge, Mozos,
2533:   Ordej\'{o}n, Taylor, and Stokbro}}]{Brandbyge:prb02}
2534: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Brandbyge}},
2535:   \bibinfo{author}{\bibfnamefont{J.~L.} \bibnamefont{Mozos}},
2536:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Ordej\'{o}n}},
2537:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Taylor}}, \bibnamefont{and}
2538:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Stokbro}},
2539:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{65}},
2540:   \bibinfo{pages}{165401} (\bibinfo{year}{2002}).
2541: 
2542: \bibitem[{\citenamefont{Drchal et~al.}(2002)\citenamefont{Drchal,
2543:   Kudrnovsk\'{y}, Bruno, Dederichs, Turek, and Weinberger}}]{Drchal:prb02}
2544: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Drchal}},
2545:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kudrnovsk\'{y}}},
2546:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Bruno}},
2547:   \bibinfo{author}{\bibfnamefont{P.~H.} \bibnamefont{Dederichs}},
2548:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Turek}}, \bibnamefont{and}
2549:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Weinberger}},
2550:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{65}},
2551:   \bibinfo{pages}{214414} (\bibinfo{year}{2002}).
2552: 
2553: \bibitem[{\citenamefont{Wortmann
2554:   et~al.}(2002{\natexlab{a}})\citenamefont{Wortmann, Ishida, and
2555:   Bl{\"{u}}gel}}]{Wortmann:prb02a}
2556: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Wortmann}},
2557:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ishida}}, \bibnamefont{and}
2558:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bl{\"{u}}gel}},
2559:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{65}},
2560:   \bibinfo{pages}{165103} (\bibinfo{year}{2002}{\natexlab{a}}).
2561: 
2562: \bibitem[{\citenamefont{Wortmann
2563:   et~al.}(2002{\natexlab{b}})\citenamefont{Wortmann, Ishida, and
2564:   Bl{\"{u}}gel}}]{Wortmann:prb02b}
2565: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Wortmann}},
2566:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ishida}}, \bibnamefont{and}
2567:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bl{\"{u}}gel}},
2568:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{66}},
2569:   \bibinfo{pages}{075113} (\bibinfo{year}{2002}{\natexlab{b}}).
2570: 
2571: \bibitem[{\citenamefont{Weinberger}(2003)}]{Weinberger:prp03}
2572: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Weinberger}},
2573:   \bibinfo{journal}{Phys. Rep.} \textbf{\bibinfo{volume}{377}},
2574:   \bibinfo{pages}{281} (\bibinfo{year}{2003}).
2575: 
2576: \bibitem[{\citenamefont{Thygesen et~al.}(2003)\citenamefont{Thygesen,
2577:   Bollinger, and Jacobsen}}]{Thygesen:prb03}
2578: \bibinfo{author}{\bibfnamefont{K.~S.} \bibnamefont{Thygesen}},
2579:   \bibinfo{author}{\bibfnamefont{M.~V.} \bibnamefont{Bollinger}},
2580:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{K.~W.}
2581:   \bibnamefont{Jacobsen}}, \bibinfo{journal}{Phys. Rev. B}
2582:   \textbf{\bibinfo{volume}{67}}, \bibinfo{pages}{115404}
2583:   (\bibinfo{year}{2003}).
2584: 
2585: \bibitem[{\citenamefont{Mavropoulos et~al.}(2004)\citenamefont{Mavropoulos,
2586:   Papanikolaou, and Dederichs}}]{Mavropoulos:prb04}
2587: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Mavropoulos}},
2588:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Papanikolaou}},
2589:   \bibnamefont{and}
2590:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Dederichs}},
2591:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{69}},
2592:   \bibinfo{pages}{125104} (\bibinfo{year}{2004}).
2593: 
2594: \bibitem[{\citenamefont{Mathon}(1997{\natexlab{a}})}]{Mathon:prb97a}
2595: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Mathon}},
2596:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{55}},
2597:   \bibinfo{pages}{960} (\bibinfo{year}{1997}{\natexlab{a}}).
2598: 
2599: \bibitem[{\citenamefont{Mathon et~al.}(1997)\citenamefont{Mathon, Umerski, and
2600:   Villeret}}]{Mathon:prb97b}
2601: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Mathon}},
2602:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Umerski}}, \bibnamefont{and}
2603:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Villeret}},
2604:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{55}},
2605:   \bibinfo{pages}{14378} (\bibinfo{year}{1997}).
2606: 
2607: \bibitem[{\citenamefont{Tsymbal and Pettifor}(1997)}]{Tsymbal:jp97}
2608: \bibinfo{author}{\bibfnamefont{E.~Y.} \bibnamefont{Tsymbal}} \bibnamefont{and}
2609:   \bibinfo{author}{\bibfnamefont{D.~G.} \bibnamefont{Pettifor}},
2610:   \bibinfo{journal}{J. Phys.: Condens. Matter.} \textbf{\bibinfo{volume}{9}},
2611:   \bibinfo{pages}{L411} (\bibinfo{year}{1997}).
2612: 
2613: \bibitem[{\citenamefont{Sanvito et~al.}(1999)\citenamefont{Sanvito, Lambert,
2614:   Jefferson, and Bratkovsky}}]{Sanvito:prb99}
2615: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sanvito}},
2616:   \bibinfo{author}{\bibfnamefont{C.~J.} \bibnamefont{Lambert}},
2617:   \bibinfo{author}{\bibfnamefont{J.~H.} \bibnamefont{Jefferson}},
2618:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.~M.}
2619:   \bibnamefont{Bratkovsky}}, \bibinfo{journal}{Phys. Rev. B}
2620:   \textbf{\bibinfo{volume}{59}}, \bibinfo{pages}{11936} (\bibinfo{year}{1999}).
2621: 
2622: \bibitem[{\citenamefont{Velev and Chang}(2003)}]{Velev:prb03}
2623: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Velev}} \bibnamefont{and}
2624:   \bibinfo{author}{\bibfnamefont{Y.~C.} \bibnamefont{Chang}},
2625:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{67}},
2626:   \bibinfo{pages}{144425} (\bibinfo{year}{2003}).
2627: 
2628: \bibitem[{\citenamefont{Velev and Butler}(2004)}]{Velev:prb04}
2629: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Velev}} \bibnamefont{and}
2630:   \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Butler}},
2631:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{69}},
2632:   \bibinfo{pages}{024404} (\bibinfo{year}{2004}).
2633: 
2634: \bibitem[{\citenamefont{Imry}(2002)}]{Imry:02}
2635: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Imry}},
2636:   \emph{\bibinfo{title}{Introduction to Mesoscopic Physics}}
2637:   (\bibinfo{publisher}{Oxford University Press}, \bibinfo{address}{Oxford},
2638:   \bibinfo{year}{2002}), \bibinfo{edition}{2nd} ed.
2639: 
2640: \bibitem[{\citenamefont{Datta}(1995)}]{Datta:95}
2641: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Datta}},
2642:   \emph{\bibinfo{title}{Electronic Transport in Mesoscopic Systems}}
2643:   (\bibinfo{publisher}{Cambridge University Press},
2644:   \bibinfo{address}{Cambridge}, \bibinfo{year}{1995}).
2645: 
2646: \bibitem[{\citenamefont{Fisher and Lee}(1981)}]{Fisher:prb81}
2647: \bibinfo{author}{\bibfnamefont{D.~S.} \bibnamefont{Fisher}} \bibnamefont{and}
2648:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
2649:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{23}},
2650:   \bibinfo{pages}{R6851} (\bibinfo{year}{1981}).
2651: 
2652: \bibitem[{\citenamefont{Beenakker}(1997)}]{Beenakker:rmp97}
2653: \bibinfo{author}{\bibfnamefont{C.~W.~J.} \bibnamefont{Beenakker}},
2654:   \bibinfo{journal}{Rev. Mod. Phys.} \textbf{\bibinfo{volume}{69}},
2655:   \bibinfo{pages}{731} (\bibinfo{year}{1997}).
2656: 
2657: \bibitem[{\citenamefont{K{\"{u}}bler}(2002)}]{Kubler:00}
2658: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{K{\"{u}}bler}},
2659:   \emph{\bibinfo{title}{Theory of Itinerant Electron Magnetism}}
2660:   (\bibinfo{publisher}{Oxford University Press}, \bibinfo{address}{Oxford},
2661:   \bibinfo{year}{2002}).
2662: 
2663: \bibitem[{fn:({\natexlab{a}})}]{fn:total_en}
2664: \bibinfo{note}{Because the magnetism of transition metals depends very
2665:   sensitively on atomic structure,\cite{Kubler:00} it is important to know this
2666:   structure quite accurately. The current drive to make devices whose lateral
2667:   dimensions approach the nanoscale means that it is becoming increasingly
2668:   important to know the atomic structures of these small systems
2669:   microscopically while at the same time it is more difficult to do this
2670:   characterization experimentally. It has become a practical alternative to
2671:   determine minimum-energy structures theoretically by minimizing as a function
2672:   of the atomic positions the total energy obtained by solving the
2673:   Schr{\"{o}}dinger equation self-consistently within the local density
2674:   approximation (LDA) of Density Functional Theory (DFT), thereby avoiding the
2675:   use of any free parameters}.
2676: 
2677: \bibitem[{\citenamefont{Bruno}(1993)}]{Bruno:jmmm93}
2678: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Bruno}}, \bibinfo{journal}{J.
2679:   Magn. \& Magn. Mater.} \textbf{\bibinfo{volume}{121}}, \bibinfo{pages}{248}
2680:   (\bibinfo{year}{1993}).
2681: 
2682: \bibitem[{\citenamefont{Stiles}(1996{\natexlab{a}})}]{Stiles:jap96}
2683: \bibinfo{author}{\bibfnamefont{M.~D.} \bibnamefont{Stiles}},
2684:   \bibinfo{journal}{J. Appl. Phys.} \textbf{\bibinfo{volume}{79}},
2685:   \bibinfo{pages}{5805} (\bibinfo{year}{1996}{\natexlab{a}}).
2686: 
2687: \bibitem[{\citenamefont{Stiles}(1996{\natexlab{b}})}]{Stiles:prb96}
2688: \bibinfo{author}{\bibfnamefont{M.~D.} \bibnamefont{Stiles}},
2689:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{54}},
2690:   \bibinfo{pages}{14679} (\bibinfo{year}{1996}{\natexlab{b}}).
2691: 
2692: \bibitem[{\citenamefont{Stiles and Penn}(2000)}]{Stiles:prb00}
2693: \bibinfo{author}{\bibfnamefont{M.~D.} \bibnamefont{Stiles}} \bibnamefont{and}
2694:   \bibinfo{author}{\bibfnamefont{D.~R.} \bibnamefont{Penn}},
2695:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{61}},
2696:   \bibinfo{pages}{3200} (\bibinfo{year}{2000}).
2697: 
2698: \bibitem[{\citenamefont{Stiles and Hamann}(1988)}]{Stiles:prb88}
2699: \bibinfo{author}{\bibfnamefont{M.~D.} \bibnamefont{Stiles}} \bibnamefont{and}
2700:   \bibinfo{author}{\bibfnamefont{D.~R.} \bibnamefont{Hamann}},
2701:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{38}},
2702:   \bibinfo{pages}{2021} (\bibinfo{year}{1988}).
2703: 
2704: \bibitem[{\citenamefont{Stiles and Hamann}(1991)}]{Stiles:prl91}
2705: \bibinfo{author}{\bibfnamefont{M.~D.} \bibnamefont{Stiles}} \bibnamefont{and}
2706:   \bibinfo{author}{\bibfnamefont{D.~R.} \bibnamefont{Hamann}},
2707:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{66}},
2708:   \bibinfo{pages}{3179} (\bibinfo{year}{1991}).
2709: 
2710: \bibitem[{\citenamefont{Schep et~al.}(1998)\citenamefont{Schep, Kelly, and
2711:   Bauer}}]{Schep:prb98}
2712: \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Schep}},
2713:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}}, \bibnamefont{and}
2714:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2715:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{57}},
2716:   \bibinfo{pages}{8907} (\bibinfo{year}{1998}).
2717: 
2718: \bibitem[{\citenamefont{van Hoof}(1997)}]{vanHoof:97}
2719: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{van Hoof}}, Ph.D. thesis,
2720:   \bibinfo{school}{University of Nijmegen}, \bibinfo{address}{Nijmegen, The
2721:   Netherlands} (\bibinfo{year}{1997}).
2722: 
2723: \bibitem[{\citenamefont{van Hoof et~al.}(1998)\citenamefont{van Hoof, Schep,
2724:   Kelly, and Bauer}}]{vanHoof:jmmm98}
2725: \bibinfo{author}{\bibfnamefont{J.~B. A.~N.} \bibnamefont{van Hoof}},
2726:   \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Schep}},
2727:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}}, \bibnamefont{and}
2728:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2729:   \bibinfo{journal}{J. Magn. \& Magn. Mater.}
2730:   \textbf{\bibinfo{volume}{177-181}}, \bibinfo{pages}{188}
2731:   (\bibinfo{year}{1998}).
2732: 
2733: \bibitem[{\citenamefont{Andersen and Jepsen}(1984)}]{Andersen:prl84}
2734: \bibinfo{author}{\bibfnamefont{O.~K.} \bibnamefont{Andersen}} \bibnamefont{and}
2735:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Jepsen}},
2736:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{53}},
2737:   \bibinfo{pages}{2571} (\bibinfo{year}{1984}).
2738: 
2739: \bibitem[{\citenamefont{Andersen et~al.}(1985)\citenamefont{Andersen, Jepsen,
2740:   and Gl{\"{o}}tzel}}]{Andersen:85}
2741: \bibinfo{author}{\bibfnamefont{O.~K.} \bibnamefont{Andersen}},
2742:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Jepsen}}, \bibnamefont{and}
2743:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Gl{\"{o}}tzel}}, in
2744:   \emph{\bibinfo{booktitle}{Highlights of Condensed Matter Theory}}, edited by
2745:   \bibinfo{editor}{\bibfnamefont{F.}~\bibnamefont{Bassani}},
2746:   \bibinfo{editor}{\bibfnamefont{F.}~\bibnamefont{Fumi}}, \bibnamefont{and}
2747:   \bibinfo{editor}{\bibfnamefont{M.~P.} \bibnamefont{Tosi}}
2748:   (\bibinfo{publisher}{North-Holland}, \bibinfo{address}{Amsterdam},
2749:   \bibinfo{year}{1985}), pp. \bibinfo{pages}{59--176}.
2750: 
2751: \bibitem[{\citenamefont{Andersen et~al.}(1986)\citenamefont{Andersen,
2752:   Pawlowska, and Jepsen}}]{Andersen:prb86}
2753: \bibinfo{author}{\bibfnamefont{O.~K.} \bibnamefont{Andersen}},
2754:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Pawlowska}},
2755:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Jepsen}},
2756:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{34}},
2757:   \bibinfo{pages}{5253} (\bibinfo{year}{1986}).
2758: 
2759: \bibitem[{\citenamefont{Andersen et~al.}(1987)\citenamefont{Andersen, Jepsen,
2760:   and Sob}}]{Andersen:87}
2761: \bibinfo{author}{\bibfnamefont{O.~K.} \bibnamefont{Andersen}},
2762:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Jepsen}}, \bibnamefont{and}
2763:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Sob}}, in
2764:   \emph{\bibinfo{booktitle}{Electronic Band Structure and its Applications}},
2765:   edited by \bibinfo{editor}{\bibfnamefont{M.}~\bibnamefont{Yussouff}}
2766:   (\bibinfo{publisher}{Springer Lecture Notes}, \bibinfo{address}{Berlin},
2767:   \bibinfo{year}{1987}), vol. \bibinfo{volume}{283}, pp.
2768:   \bibinfo{pages}{1--57}.
2769: 
2770: \bibitem[{\citenamefont{Turek et~al.}(1997)\citenamefont{Turek, Drchal,
2771:   Kudrnovsk\'{y}, \v{S}ob, and Weinberger}}]{Turek:97}
2772: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Turek}},
2773:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Drchal}},
2774:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kudrnovsk\'{y}}},
2775:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{\v{S}ob}}, \bibnamefont{and}
2776:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Weinberger}},
2777:   \emph{\bibinfo{title}{Electronic Structure of Disordered Alloys, Surfaces and
2778:   Interfaces}} (\bibinfo{publisher}{Kluwer},
2779:   \bibinfo{address}{Boston-London-Dordrecht}, \bibinfo{year}{1997}).
2780: 
2781: \bibitem[{\citenamefont{Ando}(1991)}]{Ando:prb91}
2782: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Ando}}, \bibinfo{journal}{Phys.
2783:   Rev. B} \textbf{\bibinfo{volume}{44}}, \bibinfo{pages}{8017}
2784:   (\bibinfo{year}{1991}).
2785: 
2786: \bibitem[{\citenamefont{Xia et~al.}(2002{\natexlab{a}})\citenamefont{Xia,
2787:   Kelly, Bauer, Brataas, and Turek}}]{Xia:prb02}
2788: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Xia}},
2789:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2790:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2791:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Brataas}}, \bibnamefont{and}
2792:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Turek}},
2793:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{65}},
2794:   \bibinfo{pages}{220401 (R)} (\bibinfo{year}{2002}{\natexlab{a}}).
2795: 
2796: \bibitem[{\citenamefont{Zwierzycki et~al.}(2005)\citenamefont{Zwierzycki,
2797:   Tserkovnyak, Kelly, Brataas, and Bauer}}]{Zwierzycki:prb05}
2798: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zwierzycki}},
2799:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tserkovnyak}},
2800:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2801:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Brataas}}, \bibnamefont{and}
2802:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2803:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{71}},
2804:   \bibinfo{pages}{064420} (\bibinfo{year}{2005}).
2805: 
2806: \bibitem[{\citenamefont{Xia et~al.}(2002{\natexlab{b}})\citenamefont{Xia,
2807:   Kelly, Bauer, and Turek}}]{Xia:prl02}
2808: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Xia}},
2809:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2810:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2811:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Turek}},
2812:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{89}},
2813:   \bibinfo{pages}{166603} (\bibinfo{year}{2002}{\natexlab{b}}).
2814: 
2815: \bibitem[{\citenamefont{Zwierzycki et~al.}(2003)\citenamefont{Zwierzycki, Xia,
2816:   Kelly, Bauer, and Turek}}]{Zwierzycki:prb03}
2817: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zwierzycki}},
2818:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Xia}},
2819:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2820:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2821:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Turek}},
2822:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{67}},
2823:   \bibinfo{pages}{092401} (\bibinfo{year}{2003}).
2824: 
2825: \bibitem[{\citenamefont{Bauer et~al.}(2004)\citenamefont{Bauer, Brataas,
2826:   Tserkovnyak, Halperin, Zwierzycki, and Kelly}}]{Bauer:prl04}
2827: \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
2828:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Brataas}},
2829:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tserkovnyak}},
2830:   \bibinfo{author}{\bibfnamefont{B.~I.} \bibnamefont{Halperin}},
2831:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zwierzycki}},
2832:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2833:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{92}},
2834:   \bibinfo{pages}{126601} (\bibinfo{year}{2004}).
2835: 
2836: \bibitem[{\citenamefont{Andersen and Saha-Dasgupta}(2000)}]{Andersen:prb00}
2837: \bibinfo{author}{\bibfnamefont{O.~K.} \bibnamefont{Andersen}} \bibnamefont{and}
2838:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Saha-Dasgupta}},
2839:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{62}},
2840:   \bibinfo{pages}{R16219} (\bibinfo{year}{2000}).
2841: 
2842: \bibitem[{\citenamefont{Tank and Arcangeli}(2000)}]{Tank:pssb00}
2843: \bibinfo{author}{\bibfnamefont{R.~W.} \bibnamefont{Tank}} \bibnamefont{and}
2844:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Arcangeli}},
2845:   \bibinfo{journal}{phys. stat. sol. B} \textbf{\bibinfo{volume}{217}},
2846:   \bibinfo{pages}{89} (\bibinfo{year}{2000}).
2847: 
2848: \bibitem[{\citenamefont{Andersen et~al.}(2000)\citenamefont{Andersen,
2849:   Saha-Dasgupta, Tank, Arcangeli, Jepsen, and Krier}}]{Andersen:00}
2850: \bibinfo{author}{\bibfnamefont{O.~K.} \bibnamefont{Andersen}},
2851:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Saha-Dasgupta}},
2852:   \bibinfo{author}{\bibfnamefont{R.~W.} \bibnamefont{Tank}},
2853:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Arcangeli}},
2854:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Jepsen}}, \bibnamefont{and}
2855:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Krier}}, in
2856:   \emph{\bibinfo{booktitle}{Electronic Structure and Physical Properties of
2857:   Solids: The Uses of the LMTO Method}}, edited by
2858:   \bibinfo{editor}{\bibfnamefont{H.}~\bibnamefont{Dreysse}}
2859:   (\bibinfo{publisher}{Springer Lecture Notes in Physics},
2860:   \bibinfo{address}{New York}, \bibinfo{year}{2000}), pp.
2861:   \bibinfo{pages}{3--84}.
2862: 
2863: \bibitem[{\citenamefont{Khomyakov et~al.}(2005)\citenamefont{Khomyakov, Brocks,
2864:   Karpan, Zwierzycki, and Kelly}}]{Khomyakov:prb05}
2865: \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Khomyakov}},
2866:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Brocks}},
2867:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Karpan}},
2868:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zwierzycki}},
2869:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Kelly}},
2870:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{72}},
2871:   \bibinfo{pages}{035450} (\bibinfo{year}{2005}).
2872: 
2873: \bibitem[{\citenamefont{von Barth and Hedin}(1972)}]{vonBarth:jpc72}
2874: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{von Barth}} \bibnamefont{and}
2875:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Hedin}}, \bibinfo{journal}{J.
2876:   Phys. C: Sol. State Phys.} \textbf{\bibinfo{volume}{5}},
2877:   \bibinfo{pages}{1629} (\bibinfo{year}{1972}).
2878: 
2879: \bibitem[{fn:({\natexlab{b}})}]{fn:WS_radii}
2880: \bibinfo{note}{The larger and smaller lattice constants correspond to
2881:   Wigner-Seitz atomic sphere radii of 2.669 and 2.621 Bohr atomic units,
2882:   respectively.}
2883: 
2884: \bibitem[{\citenamefont{Perdew and Zunger}(1981)}]{Perdew:prb81}
2885: \bibinfo{author}{\bibfnamefont{J.~P.} \bibnamefont{Perdew}} \bibnamefont{and}
2886:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Zunger}},
2887:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{23}},
2888:   \bibinfo{pages}{5048} (\bibinfo{year}{1981}).
2889: 
2890: \bibitem[{\citenamefont{Vosko et~al.}(1980)\citenamefont{Vosko, Wilk, and
2891:   Nusair}}]{Vosko:cjp80}
2892: \bibinfo{author}{\bibfnamefont{S.~H.} \bibnamefont{Vosko}},
2893:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Wilk}}, \bibnamefont{and}
2894:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Nusair}},
2895:   \bibinfo{journal}{Canadian Journal of Physics} \textbf{\bibinfo{volume}{58}},
2896:   \bibinfo{pages}{1200} (\bibinfo{year}{1980}).
2897: 
2898: \bibitem[{fn:({\natexlab{c}})}]{fn:standard}
2899: \bibinfo{note}{We adopt a ``standard'' configuration consisting of an
2900:   \textit{fcc} lattice with lattice constant of $a=3.614\mathring{A}$, an
2901:   \textit{spd} basis, and von Barth-Hedin's exchange-correlation potential.}
2902: 
2903: \bibitem[{\citenamefont{Soven}(1967)}]{Soven:pr67}
2904: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Soven}},
2905:   \bibinfo{journal}{Phys. Rev.} \textbf{\bibinfo{volume}{156}},
2906:   \bibinfo{pages}{809} (\bibinfo{year}{1967}).
2907: 
2908: \bibitem[{\citenamefont{Bruno et~al.}(1999)\citenamefont{Bruno, Itoh, Inoue,
2909:   and Nonoyama}}]{Bruno:jmmm99}
2910: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Bruno}},
2911:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Itoh}},
2912:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Inoue}}, \bibnamefont{and}
2913:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Nonoyama}},
2914:   \bibinfo{journal}{J. Magn. \& Magn. Mater.}
2915:   \textbf{\bibinfo{volume}{198-199}}, \bibinfo{pages}{46}
2916:   (\bibinfo{year}{1999}).
2917: 
2918: \bibitem[{\citenamefont{Henry et~al.}(1996)\citenamefont{Henry, Yang, Chiang,
2919:   Holody, Loloee, {Pratt, Jr.}, and Bass}}]{Henry:prb96}
2920: \bibinfo{author}{\bibfnamefont{L.~L.} \bibnamefont{Henry}},
2921:   \bibinfo{author}{\bibfnamefont{Q.}~\bibnamefont{Yang}},
2922:   \bibinfo{author}{\bibfnamefont{W.~C.} \bibnamefont{Chiang}},
2923:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Holody}},
2924:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Loloee}},
2925:   \bibinfo{author}{\bibfnamefont{W.~P.} \bibnamefont{{Pratt, Jr.}}},
2926:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bass}},
2927:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{54}},
2928:   \bibinfo{pages}{12336} (\bibinfo{year}{1996}).
2929: 
2930: \bibitem[{\citenamefont{{de Gronckel} et~al.}(1991)\citenamefont{{de Gronckel},
2931:   Kopinga, de~Jonge, Panissod, Schill\'{e}, and {den
2932:   Broeder}}}]{deGronckel:prb91}
2933: \bibinfo{author}{\bibfnamefont{H.~A.~M.} \bibnamefont{{de Gronckel}}},
2934:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Kopinga}},
2935:   \bibinfo{author}{\bibfnamefont{W.~J.~M.} \bibnamefont{de~Jonge}},
2936:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Panissod}},
2937:   \bibinfo{author}{\bibfnamefont{J.~P.} \bibnamefont{Schill\'{e}}},
2938:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{F.~J.~A.} \bibnamefont{{den
2939:   Broeder}}}, \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{44}},
2940:   \bibinfo{pages}{9100} (\bibinfo{year}{1991}).
2941: 
2942: \bibitem[{\citenamefont{M\'{e}ny et~al.}(1991)\citenamefont{M\'{e}ny, Panissod,
2943:   and Loloee}}]{Meny:prb92}
2944: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{M\'{e}ny}},
2945:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Panissod}}, \bibnamefont{and}
2946:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Loloee}},
2947:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{45}},
2948:   \bibinfo{pages}{12269} (\bibinfo{year}{1991}).
2949: 
2950: \bibitem[{\citenamefont{Kapusta et~al.}(1999)\citenamefont{Kapusta, Fischer,
2951:   and Sch{\"{u}}tz}}]{Kapusta:jac99}
2952: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Kapusta}},
2953:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Fischer}}, \bibnamefont{and}
2954:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sch{\"{u}}tz}},
2955:   \bibinfo{journal}{J. Alloys Compd.} \textbf{\bibinfo{volume}{286}},
2956:   \bibinfo{pages}{37} (\bibinfo{year}{1999}).
2957: 
2958: \bibitem[{fn:({\natexlab{d}})}]{fn:xia01}
2959: \bibinfo{note}{Interface disorder can {\em increase} interface transmission if
2960:   electronic structure mismatch leads to a highly reflecting clean interface.
2961:   For example, for the majority-spin electrons at an Fe/Cr (001) interface,
2962:   this is the case. There, interface disorder in the form of two layers of
2963:   50\%-50\% alloy {\em reduced} the interface resistance by a factor
2964:   three\cite{Xia:prb01}}.
2965: 
2966: \bibitem[{\citenamefont{Zhang and Levy}(1991)}]{Zhang:jap91}
2967: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Zhang}} \bibnamefont{and}
2968:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Levy}}, \bibinfo{journal}{J.
2969:   Appl. Phys.} \textbf{\bibinfo{volume}{69}}, \bibinfo{pages}{4786}
2970:   (\bibinfo{year}{1991}).
2971: 
2972: \bibitem[{\citenamefont{Lee et~al.}(1993)\citenamefont{Lee, {Pratt, Jr.}, Yang,
2973:   Holody, Loloee, Schroeder, and Bass}}]{Lee:jmmm93}
2974: \bibinfo{author}{\bibfnamefont{S.~F.} \bibnamefont{Lee}},
2975:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{{Pratt, Jr.}}},
2976:   \bibinfo{author}{\bibfnamefont{Q.}~\bibnamefont{Yang}},
2977:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Holody}},
2978:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Loloee}},
2979:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Schroeder}},
2980:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bass}},
2981:   \bibinfo{journal}{J. Magn. \& Magn. Mater.} \textbf{\bibinfo{volume}{118}},
2982:   \bibinfo{pages}{L1} (\bibinfo{year}{1993}).
2983: 
2984: \bibitem[{\citenamefont{Valet and Fert}(1993)}]{Valet:prb93}
2985: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Valet}} \bibnamefont{and}
2986:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Fert}},
2987:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{48}},
2988:   \bibinfo{pages}{7099} (\bibinfo{year}{1993}).
2989: 
2990: \bibitem[{\citenamefont{{Pratt, Jr.} et~al.}(1991)\citenamefont{{Pratt, Jr.},
2991:   Lee, Slaughter, Loloee, Schroeder, and Bass}}]{Pratt:prl91}
2992: \bibinfo{author}{\bibfnamefont{W.~P.} \bibnamefont{{Pratt, Jr.}}},
2993:   \bibinfo{author}{\bibfnamefont{S.~F.} \bibnamefont{Lee}},
2994:   \bibinfo{author}{\bibfnamefont{J.~M.} \bibnamefont{Slaughter}},
2995:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Loloee}},
2996:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Schroeder}},
2997:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bass}},
2998:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{66}},
2999:   \bibinfo{pages}{3060} (\bibinfo{year}{1991}).
3000: 
3001: \bibitem[{\citenamefont{Gijs et~al.}(1993)\citenamefont{Gijs, Lenczowski, and
3002:   Giesbers}}]{Gijs:prl93}
3003: \bibinfo{author}{\bibfnamefont{M.~A.~M.} \bibnamefont{Gijs}},
3004:   \bibinfo{author}{\bibfnamefont{S.~K.~J.} \bibnamefont{Lenczowski}},
3005:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~B.}
3006:   \bibnamefont{Giesbers}}, \bibinfo{journal}{Phys. Rev. Lett.}
3007:   \textbf{\bibinfo{volume}{70}}, \bibinfo{pages}{3343} (\bibinfo{year}{1993}).
3008: 
3009: \bibitem[{\citenamefont{Bass and {Pratt Jr.}}(1999)}]{Bass:jmmm99}
3010: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bass}} \bibnamefont{and}
3011:   \bibinfo{author}{\bibfnamefont{W.~P.} \bibnamefont{{Pratt Jr.}}},
3012:   \bibinfo{journal}{J. Magn. \& Magn. Mater.} \textbf{\bibinfo{volume}{200}},
3013:   \bibinfo{pages}{274} (\bibinfo{year}{1999}).
3014: 
3015: \bibitem[{\citenamefont{Gijs and Bauer}(1997)}]{Gijs:ap97}
3016: \bibinfo{author}{\bibfnamefont{M.~A.~M.} \bibnamefont{Gijs}} \bibnamefont{and}
3017:   \bibinfo{author}{\bibfnamefont{G.~E.~W.} \bibnamefont{Bauer}},
3018:   \bibinfo{journal}{Adv. Phys.} \textbf{\bibinfo{volume}{46}},
3019:   \bibinfo{pages}{285} (\bibinfo{year}{1997}).
3020: 
3021: \bibitem[{\citenamefont{Tsymbal}(2000)}]{Tsymbal:prb00b}
3022: \bibinfo{author}{\bibfnamefont{E.~Y.} \bibnamefont{Tsymbal}},
3023:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{62}},
3024:   \bibinfo{pages}{R3608} (\bibinfo{year}{2000}).
3025: 
3026: \bibitem[{\citenamefont{Shpiro and Levy}(2000)}]{Shpiro:prb00}
3027: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Shpiro}} \bibnamefont{and}
3028:   \bibinfo{author}{\bibfnamefont{P.~M.} \bibnamefont{Levy}},
3029:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{63}},
3030:   \bibinfo{pages}{014419} (\bibinfo{year}{2000}).
3031: 
3032: \bibitem[{\citenamefont{Gerritsen}(2002)}]{Gerritsen:02}
3033: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Gerritsen}}, Master's thesis,
3034:   \bibinfo{school}{University of Twente}, \bibinfo{address}{Enschede, The
3035:   Netherlands} (\bibinfo{year}{2002}).
3036: 
3037: \bibitem[{\citenamefont{Khomyakov and Brocks}(2004)}]{Khomyakov:prb04}
3038: \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Khomyakov}}
3039:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Brocks}},
3040:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{70}},
3041:   \bibinfo{pages}{195402} (\bibinfo{year}{2004}).
3042: 
3043: \bibitem[{\citenamefont{Mathon}(1997{\natexlab{b}})}]{Mathon:prb97c}
3044: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Mathon}},
3045:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{56}},
3046:   \bibinfo{pages}{11810} (\bibinfo{year}{1997}{\natexlab{b}}).
3047: 
3048: \bibitem[{\citenamefont{Butler et~al.}(2001)\citenamefont{Butler, Zhang,
3049:   Schulthess, and MacLaren}}]{Butler:prb01}
3050: \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Butler}},
3051:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Zhang}},
3052:   \bibinfo{author}{\bibfnamefont{T.~C.} \bibnamefont{Schulthess}},
3053:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~M.}
3054:   \bibnamefont{MacLaren}}, \bibinfo{journal}{Phys. Rev. B}
3055:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{054416}
3056:   (\bibinfo{year}{2001}).
3057: 
3058: \bibitem[{\citenamefont{Mathon and Umerski}(2001)}]{Mathon:prb01}
3059: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Mathon}} \bibnamefont{and}
3060:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Umerski}},
3061:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{63}},
3062:   \bibinfo{pages}{220403 (R)} (\bibinfo{year}{2001}).
3063: 
3064: \bibitem[{\citenamefont{Mathon and Umerski}(2005)}]{Mathon:prb05}
3065: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Mathon}} \bibnamefont{and}
3066:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Umerski}},
3067:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{71}},
3068:   \bibinfo{pages}{220402 (R)} (\bibinfo{year}{2005}).
3069: 
3070: \bibitem[{\citenamefont{Caroli et~al.}(1971)\citenamefont{Caroli, Combescot,
3071:   Nozi\`{e}res, and Saint-James}}]{Caroli:jpc71a}
3072: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Caroli}},
3073:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Combescot}},
3074:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Nozi\`{e}res}},
3075:   \bibnamefont{and}
3076:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Saint-James}},
3077:   \bibinfo{journal}{J. Phys. C: Sol. State Phys.} \textbf{\bibinfo{volume}{4}},
3078:   \bibinfo{pages}{916} (\bibinfo{year}{1971}).
3079: 
3080: \bibitem[{\citenamefont{Krsti{\'{c}} et~al.}(2002)\citenamefont{Krsti{\'{c}},
3081:   Zhang, and Butler}}]{Krstic:prb02}
3082: \bibinfo{author}{\bibfnamefont{P.~S.} \bibnamefont{Krsti{\'{c}}}},
3083:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Zhang}}, \bibnamefont{and}
3084:   \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Butler}},
3085:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{66}},
3086:   \bibinfo{pages}{205319} (\bibinfo{year}{2002}).
3087: 
3088: \bibitem[{\citenamefont{Inglesfield}(1981)}]{Inglesfield:jpc81}
3089: \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{Inglesfield}},
3090:   \bibinfo{journal}{J. Phys. C: Sol. State Phys.}
3091:   \textbf{\bibinfo{volume}{14}}, \bibinfo{pages}{3795} (\bibinfo{year}{1981}).
3092: 
3093: \bibitem[{\citenamefont{Crampin et~al.}(1992)\citenamefont{Crampin, van Hoof,
3094:   Nekovee, and Inglesfield}}]{Crampin:jp92}
3095: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Crampin}},
3096:   \bibinfo{author}{\bibfnamefont{J.~B. A.~N.} \bibnamefont{van Hoof}},
3097:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Nekovee}}, \bibnamefont{and}
3098:   \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{Inglesfield}},
3099:   \bibinfo{journal}{J. Phys.: Condens. Matter.} \textbf{\bibinfo{volume}{4}},
3100:   \bibinfo{pages}{1475} (\bibinfo{year}{1992}).
3101: 
3102: \end{thebibliography}
3103: 
3104: \end{document}