cond-mat0402607/ssf.tex
1: \documentclass[aps,twocolumn]{revtex4}
2: 
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: 
6: \usepackage{graphicx}
7: \usepackage{color}
8: \usepackage{amsmath}
9: \usepackage{bm}
10: 
11: \begin{document}
12: \newcommand{\bea}{\begin{eqnarray}}
13: \newcommand{\eea}{\end{eqnarray}}
14: \newcommand{\be}{\begin{equation}}
15: \newcommand{\ee}{\end{equation}}
16: \newcommand{\bi}{\bibitem}
17: \newcommand{\D}{\mbox{$\Delta$}}
18: \newcommand{\mitD}{\mbox{${\mit \Delta}$}}
19: \renewcommand{\d}{\mbox{$\delta$}}
20: \renewcommand{\t}{\mbox{$\theta$}}
21: \renewcommand{\o}{\mbox{$\omega$}}
22: \renewcommand{\a}{\mbox{$\alpha$}}
23: \newcommand{\s}{\sigma}
24: \newcommand{\g}{\mbox{$\gamma$}}
25: \newcommand{\ep}{\mbox{$\epsilon$}}
26: \newcommand{\p}{\mbox{$\psi$}}
27: \newcommand{\bfk}{\mbox{${\bf k}$}}
28: \newcommand{\bfR}{\mbox{${{\bf R}}$}}
29: \newcommand{\bfr}{\mbox{${\bf r}$}}
30: \newcommand{\naR}{\mbox{$\nabla_{\bf R}$}}
31: \newcommand{\xipa}{\mbox{$\xi_{\parallel}$}}
32: \newcommand{\xipe}{\mbox{$\xi_{\perp}$}}
33: \newcommand{\xia}{\mbox{$\xi_{ani}$}}
34: \newcommand{\rr}{{\bf r}}
35: \newcommand{\kk}{{\bf k}}
36: \newcommand{\pp}{{\bf p}}
37: \newcommand{\qq}{{\bf q}}
38: \newcommand{\BSCCO}{{Bi$_2$Sr$_2$CaCu$_2$O$_8$ }}
39: \newcommand{\YBCO}{{YBa$_2$Cu$_3$O$_{7-\delta}$ }}
40: \newcommand{\nn}{\nonumber}
41: \newcommand{\G}{{\cal G}}
42: \newcommand{\bom}{{\bar\omega}}
43: \newcommand{\KK}{{\cal K} }
44: \newcommand{\EE}{{\cal E} }
45: \newcommand{\sgn}{{\rm sgn}}
46: \newcommand{\uG}{\underline{G}}
47: \newcommand{\ucG}{\underline{{\cal G}}}
48: \newcommand{\uGo}{\underline{G}^0}
49: \newcommand{\ug}{\underline{g}}
50: \newcommand{\uL}{\underline{L}}
51: \newcommand{\uM}{\underline{M}}
52: \newcommand{\uT}{\underline{T}}
53: \newcommand{\ut}{\underline{t}}
54: \newcommand{\uV}{\underline{V}}
55: \newcommand{\uVm}{\underline{V_1}}
56: \newcommand{\uU}{\underline{U}}
57: \newcommand{\uR}{\underline{R}}
58: \def\k{{\bf k}}
59: \def\r{{\bf r}}
60: \def\R{{\bf R}}
61: \def\p{{\bf p}}
62: \def\v{{\bf v}}
63: \def\D{{\cal D}}
64: %\twocolumn[\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse%
65: %\endcsname
66: 
67: \title{Electronic structure of $d$-wave superconducting quantum wires }
68: 
69: \author{A. M. Bobkov$^{1,2}$, L.-Y. Zhu$^{3}$, S.-W. Tsai$^{3, 4}$,
70: T. S. Nunner$^{3}$, Yu. S. Barash$^{1,2}$ and P. J. Hirschfeld$^{3}$\\~}
71: 
72: \affiliation{$^1$Lebedev Physical Institute, Leninsky Prospect 53,
73: Moscow 119991, Russia\\$^2$Institute of Solid State Physics,
74: Chernogolovka, Moscow reg., 142432 Russia \\$^3$Department of
75: Physics, University of Florida, Gainesville, FL
76: 32611\\$^4$Department of Physics, Boston University, Boston, MA
77: 02215}
78: 
79: \date{\today}
80: 
81: \begin{abstract}
82: We present analytical and numerical results for the electronic
83: spectra of wires of a d-wave superconductor on a square lattice.
84: The spectra of Andreev and other quasiparticle states, as well as
85: the spatial and particle-hole structures of their wave functions,
86: depend on interference effects caused by the presence of the
87: surfaces and are qualitatively different for half-filled wires
88: with even or odd number of chains. For half-filled  wires with an
89: odd number of chains $N$  at (110) orientation, spectra consist of
90: $N$ doubly degenerate branches. By contrast, for even $N$ wires,
91: these levels are split, and all quasiparticle states, even the
92: ones lying above the maximal gap, have the characteristic
93: properties of Andreev bound states. These Andreev states above the
94: gap can be interpreted as a consequence of
95: an infinite sequence of Andreev reflections experienced by
96: quasiparticles along their trajectories bounded by the surfaces of
97: the wire. Our  microscopic results for the local density of states
98: display atomic-scale Friedel oscillations due to the presence of
99: the surfaces, which should be observable by scanning tunneling
100: microscopy. For narrow wires the self-consistent treatment of the
101: order parameter is found to play a crucial role.  In particular,
102: we find that for small wire widths the finite geometry may drive
103: strong fluctuations or even stablilize exotic quasi-1D pair states
104: with spin triplet character.
105: \end{abstract}
106: 
107: %\pacs{74.25.Bt,74.25.Jb,74.40.+k}
108: 
109: \maketitle
110: 
111: \section{Introduction}
112: Since the discovery of high temperature superconductivity (HTS), the
113: origin of the pairing phenomenon in these materials has been the
114: subject of intense debate, and is still not clarified. Part of the
115: unusual nature of HTS which has hindered theoretical analysis is the
116: short coherence length, which allows short-wavelength fluctuations of
117: various types of local order to coexist with superconductivity. Most
118: probes of the nature of the superconducting state have been
119: restricted, until fairly recently, to measurements of bulk properties
120: and tunnelling through relatively large areas. Following the
121: pioneering work of Hess et al.\cite{Hess91}, it was realized that
122: scanning tunneling microscopy (STM) could provide an atomic-scale
123: picture of the superconducting state, particularly useful when
124: applied to inhomogeneous situations like the vortex lattice.
125: Measurements of this type were subsequently performed on
126: high-temperature superconductors \cite{Fischer95}. In the past few
127: years, scanning tunnelling microscopy on the surface of HTS have
128: compiled a novel and fascinating picture of the local electronic
129: structure of a few of these materials
130: \cite{yazdani,davisnative,davisZn,cren,davisinhom1,delozanne,howald,davisinhom2}.
131: In the Ba$_2$Sr$_2$CaCu$_2$O$_8$ system (BSCCO), one dramatic
132: implication of these experiments is that even relatively high
133: quality single crystals display inhomogeneous electronic structure
134: at the nanoscale \cite{davisinhom1,howald,delozanne,davisinhom2}.
135: 
136: In parallel to studies of the HTS materials, point contact
137: spectroscopy has been used to study the electronic structure of
138: ultrasmall  conventional superconducting
139: islands \cite{vonDelftreview}. Among many fascinating consequences
140: of the nanoscale geometry are number parity effects, in which the
141: qualitative electronic structure depends sensitively on whether
142: the number of electrons on the island are even or odd. More
143: recently,  superconducting wires of widths tens of
144: nanometers \cite{Tinkham} have also been fabricated.  Although
145: they have not yet been studied by STM or similar methods, this
146: should be technically feasible.
147: 
148: Isolated nanoscale grains and wires of HTS material have not been
149: fabricated to our knowledge. While this may prove technically
150: quite difficult to achieve due to the complexity of the crystal structure,
151: there seems to be no fundamental obstacle in the
152: long run.  This applies as well to other superconductors thought
153: to manifest unconventional superconducting order, where
154: effects of finite geometry should be easier to see since
155: coherence lengths tend to be larger.
156: 
157: It is our purpose in this paper to study how $d$-wave (e.g., HTS)
158: and other unconventional symmetry superconductors behave in finite
159: geometry at the atomic scale. Ziegler et al.\cite{zieglerwire}
160: began the study of this problem in the case of (100) $d$-wave
161: quantum wires, pointing out the dependence of the Fermi level
162: density of states on the parity of the wire width.   This is a
163: natural consequence of the discretization of the electronic energy
164: levels due to the finite wire width in the $d$-wave state.
165: While in the $s$-wave case all the interesting physics is tuned by
166: the level discretization, one expects a priori one  fundamental
167: difference in the $d$-wave case: for any geometry with surfaces
168: making an arbitrary angle with the crystal axes, pair-breaking
169: processes take place on a scale of the coherence length.
170: The most important consequence for the electronic structure should
171: be the formation of Andreev surface states.
172: 
173: However, little is known about how Andreev surface states behave
174: when the size of the superconductor becomes comparable to $\xi_0$.
175: The zero-energy states form on surfaces with orientations different
176: from the antinodal directions of a $d$-wave superconductor, due to
177: the sign change of the order parameter. In high-temperature superconductors,
178: such states manifest themselves as the zero-bias conductance peak
179: (ZBCP) in tunneling spectroscopy in the ab-plane \cite{geerk88,hu94,tan95,%
180: mats95,buch95,tan952,cov96,xu96,fog97,bbs97,cov97,alf97,ek97,ap98,alf981,%
181: alf982,sin98,wei98,apr99,deutsch99,Ting110,cov00,pairor02,greene02,wumou03,%
182: greene03}, the anomalous temperature behavior of the Josephson critical
183: current~\cite{tan96,bbr96,ilichev01,blamire03} and the upturn in the
184: temperature dependence of the magnetic penetration depth~\cite{walter98,%
185: bkk00,carr01} (see also review articles \cite{tan00,wendin01}).
186: The conventional description of Andreev surface states, as well as
187: the Andreev reflection itself, is based on the quasiclassical
188: approximation, a powerful tool in the theoretical study of various
189: properties of inhomogeneous superconducting systems.
190: 
191: The quasiclassical theory of superconductivity gives a so-called
192: coarse-grained description of the phenomena, averaged over
193: interatomic distances. This has been used, for example, to
194: calculate the local quasiparticle density of states (LDOS).
195: However, these coarse-grained averaged results are not adequate to
196: analyze atomic resolution measurements by STM and some other
197: contemporary experimental techniques (e.g. atomic force microscopy
198: \cite{giessibl}). To obtain this type of information, a fully
199: quantum mechanical atomic-scale approach going beyond the
200: quasiclassical approximation in describing inhomogeneous states of
201: superconductors is required. We address this problem in the
202: present paper using a tight-binding BCS-like model of $d$-wave
203: superconductor on a square lattice.
204: 
205: As a first step towards understanding the effect of constrained
206: geometry, we study the simplest case of $d$-wave superconducting
207: wires  consisting of $N$ parallel chains as the system size is
208: reduced. The  quasiparticle spectrum of such systems is described
209: both analytically, with an assumed spatially homogeneous order
210: parameter, and also numerically with a fully self-consistent
211: approach. In the limit $N\gg 1$ the usual surface Andreev states
212: in lattice models \cite{Ting110,tan00,pairor02}, as well as the
213: surface states known in continuous models \cite{bbs97}, can be
214: recovered at each surface of the wire. However, for sufficiently
215: narrow wires, when the transverse wire dimension is the order of
216: the superconducting coherence length, the Andreev states strongly
217: interfere and give rise to qualitatively new effects. We show
218: below that only those effects which occur for bands sufficiently
219: far from half-filling and relatively wide wires  can be described
220: with the quasiclassical theory of superconductivity. In addition,
221: we demonstrate how and under what  conditions one can recover
222: earlier quasiclassical results for Andreev states in $d$-wave
223: superconducting films \cite{nagai} from our microscopic approach.
224: 
225: The microscopic method adopted in this paper to constrain the
226:  geometry involves introducing lines of impurities of
227: potential strength taken to infinity to bound the wire.
228:  We  show that earlier microscopic results on (110) surfaces
229:  \cite{bbs97,Ting110,pairor02,Walker,tan00} and (100) wires
230: \cite{zieglerwire} can be reproduced by this technique, and then
231:  extend it to calculate new results on  wires with other
232:  orientations.  We find that the results for electronic spectra
233:  are very sensitive to the number parity of the wire width, and
234:  that true zero-energy Andreev states can only exist in wires with
235:  odd numbers of chains.  In even wires, the Andreev states are
236:  split,
237:  pushed away from the Fermi energy, and can have either surface or
238:  standing wave character.   Finally, we show that for smaller
239:  wires self-consistency effects become important and can even,
240:  within mean field theory, lead to condensation of a fully gapped
241:  spin triplet state instead of $d$-wave order.
242: 
243: Our results may also have some qualitative relevance for the
244: related problem of $d$-wave superconducting grains, where the
245: geometry is constrained in 2 dimensions.  In addition to
246: artificially fabricated islands, some authors have proposed that
247: the BSCCO-2212 samples which display nanoscale inhomogeneity
248: should be thought of as a collection of weakly coupled $d$-wave
249: grains of roughly the size of the superconducting coherence length
250: $\xi_0$, or $d$-wave grains coupled to grains of another
251: electronic phase \cite{davisinhom1,howald,davisinhom2,Joglekar}.
252: In fact, the structure of a general, possibly irregular small
253: grain of $d$-wave superconductor has not been studied to our
254: knowledge. Understanding how the local density of states (LDOS) of
255: these wires depend on the wire width and the orientation, as well
256: as on the deviation from half-filling, could provide important
257: intuition for the question of the electronic structure of the
258: small irregular grains possibly present in BSCCO samples.
259: 
260: The outline of the paper is as follows. In Sec.\ref{sec:Tmatrix},
261: we introduce the formalism for the problem.  In
262: Sec.\ref{sec:surface}, we discuss three special semi-infinite
263: surface orientations: $(110)$, $(210)$ and $(100)$. In Sec.\ref
264: {sec:swires}, we study the $(110)$ superconducting wires with even
265: and odd width and make an comparison with the discrete states in
266: normal metal wires (Sec.\ref{sec:nwires}) to try to identify the
267: nature of the true Andreev states. In Sec.\ref{sec:finitemu}, we
268: study the effects of deviations from the half-filling. In
269: Sec.\ref{sec:self} the results of fully self-consistent
270: calculations are presented. In particular, we allow the order
271: parameter to vary spatially and comment on the differences in our
272: results. In the case of narrow wires the self-consistent study
273: shows the appearance of some peculiar types of superconducting
274: pairing. Finally, in Sec.\ref{sec:conclusion}, we present our
275: conclusions.
276: 
277: \section{Model description and formalism}
278: \label{sec:Tmatrix}
279: The Hamiltonian for a pure singlet superconductor
280: can be written as:
281: \begin{eqnarray}
282: \cal{H} &=& - t \sum_{\langle i,j\rangle, \sigma}
283: c^\dagger_{i\sigma} c_{j\sigma} - \sum_{i,\sigma} [\mu - U_i]
284: c^\dagger_{i\sigma} c_{i\sigma} \nonumber \\ &+& \sum_{\langle
285: i,j\rangle} \{ \Delta_{ij} c^\dagger_{i\uparrow}
286: c^\dagger_{j\downarrow} + h.c. \}, \label{ham}
287: \end{eqnarray}
288: where we have chosen a  nearest neighbor tight-binding band for
289: simplicity; $\mu$ is the chemical potential.
290: A superconducting pairing is defined for nearest neighbors
291: $\Delta_{ij} = -V \langle c_{j\downarrow}c_{i\uparrow}
292: \rangle$ on the bond $\{i,j\}$. The parameter $t$ is of order 150
293: meV for high $T_c$ materials, and we consider both the
294: particle-hole symmetric model $\mu=0$ and the more realistic case
295: of finite $\mu$. In non-self-consistent calculations the OP has
296: the familiar $k$-space form $\Delta_\k = \Delta_0 [ \cos(k_aa) -
297: \cos(k_ba) ]$, where $ \Delta_0 = \frac{1}{2}\sum_\pm (
298: \Delta_{i\,i\pm r_a} - \Delta_{i\,i\pm r_b} ) $ is independent of
299: $i$, and  is taken to be $0.2t$. The lattice constant is denoted by
300: $a$. The maximum gap is $\Delta_{max}=2\Delta_0$. We also present
301: self-consistent calculations, in which case $V$ is chosen to yield
302: this same value of $\Delta_0$ far from wire edges.
303: 
304: It is possible to constrain the geometry underlying Eq.
305: (\ref{ham}) in several different ways. We present results here for
306: a method discussed, for example, in Refs.\
307: \onlinecite{MorrDemler,Ting110} in which the on-site potentials
308: $U_i$ are chosen to lie on the boundary and their value is taken
309: to infinity to cut off electron transport through the boundary.
310: This technique has the virtue that the strength of the barrier can
311: in principle be lowered to allow different degrees of transparency
312: and the study of tunneling phenomena. In this work we restrict
313: ourselves to impurity configurations and strengths for which the
314: constrained system is completely isolated from its environment.
315: This is equivalent to the assumption of open boundaries, when no
316: hopping and no pairing take place outside the region. No disorder
317: is introduced in the system and we consider $U_i$ as ``impurity''
318: potentials only so as to form the surfaces of the superconducting
319: region.
320: 
321: The Hamiltonian of the surface term is taken to be \be U
322: =U_0\sum_{\ell} c_\ell^\dagger c_\ell, \ee where the set of sites
323: $\ell$ is determined exclusively by the boundaries of the desired
324: system (see below). The full Fourier space Green's function for
325: the system in the presence of these impurities is quite generally:
326: \begin{eqnarray}
327: \check{G}(\k,\k^{\prime},\omega) &=& \check{G}^{(0)}(\k,\omega)
328: \delta_{\k,\k^{\prime}} +\nonumber\\
329: &+&\check{G}^{(0)}(\k,\omega)
330: \check{T}(\k,\k^{\prime},\omega) \check{G}^{(0)}(\k^{\prime},\omega)
331: \enspace ,
332: \label{greens}
333: \end{eqnarray}
334: where the $T$-matrix can be found from the following equations:
335: \begin{eqnarray}
336: \check{T}(\k,\k^{\prime},\omega)&=&\check{U}(\k,\k^{\prime})+\nonumber\\
337: &+&\sum_{\k^{''}} \check{U}(\k,\k^{''})
338: \check{G}^{(0)}(\k^{''},\omega)
339: \check{T}(\k^{''},\k^{\prime},\omega) . \label{tmat}
340: \end{eqnarray}
341: Here $\check{G}$ and $\check{T}$ take $4\times4$ matrix form in
342: the four-dimensional product space of particle-hole and spin
343: variables. If we choose nonmagnetic on-site potentials and
344: consider singlet superconductors, the problem reduces to $2\times
345: 2$ matrices in Nambu space. The Nambu retarded propagator for the
346: pure $d$-wave superconductor is
347: 
348: \begin{equation}
349: \hat{G}^{(0)}(\k,\omega)={\omega \hat{\tau}_0 +\xi_\k \hat{\tau}_3
350: + \Delta_\k\hat{\tau}_1 \over (\omega+i0)^2 -
351: \xi_\k^2-\Delta_\k^2} \enspace ,
352: \end{equation}
353: where the $\tau_\alpha$ are the Pauli matrices and
354: $\xi_{\bf k}=-2t[\cos(k_aa)+\cos(k_ba)]-\mu$.
355: Calculating the  $T$-matrix allows us to  obtain the
356: eigenenergies of the system from the condition
357: ${\rm det }\hat{T}^{-1}(\omega)=0$.
358: 
359: The local spin-resolved quasiparticle density of states is given as
360: \begin{eqnarray}
361:     \rho_{\uparrow(\downarrow)}({\bf r},\omega) &=& -\pi^{-1}\mbox{Im }
362:     G_{11,\uparrow\uparrow(\downarrow\downarrow)}({\bfr},\r,
363: \omega) \enspace .
364: \end{eqnarray}
365: After integration of the LDOS over energy, we should
366: obtain the total number of quasiparticle states per one site.
367: Since each site on the lattice possesses 2 states with opposite
368: spins, the spin-resolved LDOS normalization is
369: 
370: \begin{equation}
371: \int_{-\infty}^\infty d\omega \rho_{\uparrow (\downarrow)}(\r,\omega) = 1.
372: \end{equation}
373: 
374: 
375: \section{surface case }
376: 
377: \label{sec:surface}
378: 
379: Upon a conventional reflection on the (110) surface of a $d$-wave
380: superconductor, the order parameter always changes sign as the
381: direction of the quasiparticle momentum $\k$ is varied. This leads
382: to  Andreev reflection and, eventually, to the formation of the
383: dispersionless zero-energy surface Andreev bound states. For other
384: surface orientations the sign change does not take place for all
385: incoming momentum directions.
386:  It is important to notice that the number of consecutive
387: impurity lines that are needed in order to cut the system depends
388: on the orientation of the surface one wants to consider. For (100)
389: and (110) surfaces and nearest neighbor hopping, one line of
390: impurities is sufficient to cut communication between the two
391: sides. For a (210) surface, the nearest neighbor hopping and
392: pairing terms can still connect this particle with another across
393: a single impurity line, so for a simple tight-binding band,  two
394: consecutive lines are needed to close the system (Fig.
395: \ref{fig:surf210}). Alternatively, if one includes a next-nearest
396: neighbor hopping $t'$, even a (110) surface is not closed by a
397: single line of infinite impurities; the system considered in Ref.
398: \cite{MorrDemler} is therefore not a closed (impenetrable)
399: surface. For a ($hl0$) surface in general, max({$h,l$}) lines are
400: needed for a model that includes nearest-neighbor terms only, and
401: $h+l$ lines are needed for a model that includes
402: next-nearest-neighbor terms. Clearly, the technique becomes
403: cumbersome for arbitrary angles, but one can nevertheless learn a
404: good deal by considering special cases.
405: 
406: 
407: \begin{figure}[tbh]
408: \begin{center}
409: \leavevmode
410: \includegraphics[width=.6\columnwidth]{surface1.eps}
411: \caption{(210) surface or wire. Two lines of impurities (closed
412: squares) are needed to isolate the lattice sites (closed circles)
413: with  nearest neighbor hopping.} \label{fig:surf210}
414: \end{center}
415: \end{figure}
416: 
417: 
418: 
419:  In the presence of a surface of arbitrary orientation, the
420: simplest way of applying Bloch's theorem to this discrete system
421: is by using a surface-adapted Brillouin zone\cite{pairor02,Walker}.
422: We define new coordinates ($\hat{x}$, $\hat{y}$), rotated with
423: respect to the crystal axes ($\hat{a}$, $\hat{b}$), where
424: $\hat{x}$ is the direction normal to the surface and $\hat{y}$ is
425: the direction along the surface. The system is periodic
426: along the $y$-direction and the crystal momentum component $k_y$
427: of a quasiparticle is conserved. Instead of the usual
428: square Brillouin zone $k_a = [-\pi, \pi]$, $k_b = [-\pi, \pi]$
429: (for unit lattice constant $a=1$) we now use the surface-adapted
430: Brillouin zone given by $k_x =[-\pi/d, \pi/d]$ and
431: $k_y = [-\pi d, \pi d]$. Here $d =1/\sqrt{h^2+l^2}$ is the distance
432: between the nearest chains (layers) aligned along the surfaces.
433: The momenta in the two coordinate systems are simply related
434: through rotation of an angle $\theta = \tan^{-1} h/l$.
435: 
436: Now we turn to the solution of the equation for the $T$-matrix
437: (Eq. \ref{tmat}) for the case of one line of impurities. We start
438: with the ansatz
439: \begin{eqnarray}
440: \hat{T} = T_0 \hat{\tau}_0 + T_1 \hat{\tau}_1 + T_3 \hat{\tau}_3
441: \end{eqnarray}
442: and find, for arbitrary strength of impurity potential,
443: \bea T_0(k_y, \omega) &=& {{G}_0^{(0)}(0, k_y, \omega) \over
444: D_1} \ , \\
445: T_1(k_y, \omega) &=& {-{G}_1^{(0)}(0, k_y, \omega) \over D_1} \ , \\
446: T_3(k_y, \omega) &=& {c-{G}_3^{(0)}(0, k_y, \omega) \over D_1}
447: \eea
448: where $D_1(k_y, \omega) = [c-{G}_3^{(0)}(0, k_y, \omega)]^2 -
449: {G}_0^{(0)}(0, k_y, \omega)^2 + {G}_1^{(0)}(0, k_y, \omega)^2$, $c
450: = 1/V_0$  and ${G}_i^{(0)}(x, k_y, \omega)$ is the Fourier
451: transform with respect to $k_x$ of the $i$-th Nambu component of
452: the bare Green's function $G_i^{(0)}(\k, \omega)$
453: \bea
454: {G}_i^{(0)}(n, k_y, \omega) =\frac{d}{2\pi}
455: \!\!\int\limits_{k_x=-\pi/d}^{\pi/d}\!\! G^{(0)}_i (k_x, k_y, \omega) e^{i
456: k_x n d} dk_x\ .
457: \label{gtilde}
458: \eea
459: The site index $n$
460: corresponds to $x$-coordinate $x= nd$. For the case of infinitely
461: strong impurity potential considered here, $c = 0$. In this case
462: the expression for ${\hat T}$ can be written in the compact form \bea
463: \hat{T}(k_y, \omega) = - \left[\hat{{G}}^{(0)}(0, k_y, \omega)
464: \right]^{-1} \label{t} \eea and the poles of the $T$-matrix
465: correspond to zeros of the determinant of the Green's function
466: $\hat{{G}}^{(0)}(0, k_y, \omega)$.
467: 
468: The Fourier transform with respect to $k_x$ of Eq. \ref{greens} is
469: \bea &&\hat{{G}}(n, n^{\prime}, k_y, \omega) = \hat{{G}}^{(0)} (n
470: - n^{\prime}, k_y, \omega) -
471: \\ \nonumber
472: &&\ \ \ \ \ \ \  \ \ \hat{{G}}^{(0)}(n, k_y, \omega)
473: \left[\hat{{G}}^{(0)}(0, k_y, \omega)\right]^{-1}
474: \hat{{G}}^{(0)}(-n^{\prime}, k_y, \omega) \label{gn} \eea Due to
475: periodicity of the system along the $y$-direction, calculation
476: of the LDOS at site $n$ will simply involve a sum of ${G}(n,n,
477: k_y, \omega)$ over all values of $k_y$ within the surface-adapted
478: Brillouin zone and over two spin directions:
479: \begin{equation}
480: \rho(\r,\omega)=\rho(n,\omega)= -\frac{2}{\pi}{\rm Im} \int_{-\pi
481: d}^{\pi d}\frac{dk_y}{2\pi d} {G}_{11}(n,n,k_y,\omega)
482: \label{11}
483: \end{equation}
484: 
485: \subsection{(100) surface}
486: 
487: Since (100) surfaces are not pairbreaking in $d$-wave
488: superconductors, we do not {\it a priori} expect to see
489: interesting physics arising from Andreev states. On the other
490: hand, the mere presence of a surface can induce surface Tamm bands
491: \cite{Tamm}, decaying in the bulk on the atomic scale and experiencing
492: the Friedel-like oscillations of the LDOS. In our model Tamm states
493: have nothing to do with the superconductivity, although the
494: possibility for pairing of electrons occupying these surface
495: states is not excluded \cite{ginzburg}. While it is not our intent
496: to study these in detail, we present some results to show that
497: such states can be seen by the STM even in situations when surface
498: Andreev states are absent.  Fig.\ref{fig:surf100}  shows on upper
499: and lower panels the  LDOS for various distances from the surface
500: in the normal metal and the superconducting
501: states, respectively.
502: 
503: \begin{figure}[tbh]
504: \begin{center}
505: \leavevmode
506: \includegraphics[clip=true,width=.9\columnwidth]{surface100_N.eps}
507: \includegraphics[clip=true,width=0.9\columnwidth]{surface100_S.eps}
508: \caption{ Local density of states $\rho(x,\omega)$ for a (100)
509: surface. Normal state (upper panel) and superconducting state (lower panel),
510: $\rho(x,\omega)$ vs. $\omega/\Delta_{max}$ for various
511: distances $n=x/d$ from surface; $\Delta_{max}=0.4t$, $\mu=0$.}
512: \label{fig:surf100}
513: \end{center}
514: \end{figure}
515: 
516: 
517: \subsection{(110) surface}
518: 
519: Results for a (110) surface are expected to be qualitatively
520: different from the ones obtained for a (100) surface. The bulk
521: order parameter in the coordinate system of the crystal axes $\hat
522: a$ and $\hat b$ is $\Delta_\k =\Delta_0(\cos k_x a -\cos k_y a)$,
523: but in the coordinates of the surface $\hat x$ (perpendicular to
524: surface) and $\hat y$ (parallel to surface) becomes $\Delta_\k = 2
525: \Delta_0 \sin k_x d \sin k_y d$, with $d=a/\sqrt{2}$.
526:  In this
527: case, an incident particle with any non-zero $k_y$ experiences a
528: sign change in the order parameter as it reflects from the
529: surface.
530: 
531: \begin{figure}[tbh]
532: \begin{center}
533: \leavevmode
534: \includegraphics[clip=true,width=.9\columnwidth]{surface110_S.eps}
535: \caption{Local density of states $\rho(x,\omega)$ in
536: superconducting state versus $\omega/\Delta_{max}$ for the case of
537: one (110) surface, $\Delta_0=0.2t$, $\mu=0$. Left: chains an {\it
538: odd} distance $n=x/d$ from surface at $x=0$. Right: chains an {\it
539: even} distance from $x=0$. } \label{fig:surf110}
540: \end{center}
541: \end{figure}
542: 
543: Calculation of the bare Green's function $\hat{{G}}^{(0)}(n, k_y,
544: \omega)$ from Eq. \ref{gtilde} gives at $\mu=0$, $G_{1,3}(n, k_y,
545: \omega)=0$ for even $n$ and $G_{0}(n, k_y, \omega) = 0$ for odd
546: $n$. Explicitly,
547: \begin{eqnarray}
548: \hat{{G}}^{(0)}(n=2m,k_y,\omega)=\qquad\qquad\qquad\qquad\qquad\qquad
549: &\nonumber
550: \\
551: \qquad\qquad -\frac{i|\omega|\exp\left\{-i|n|z\
552: \mathrm{sgn}\left[\omega(q^2-\Delta^2)\right]\right\}}
553: {\sqrt{(\omega^2-\Delta^2)(q^2-\omega^2)}}\hat{\tau_0}
554: \label{geven}
555: \end{eqnarray}
556: and \bea \hat{{G}}^{(0)}(n=2m+1,k_y,\omega)=\qquad \qquad
557: \qquad \qquad \qquad \qquad\nonumber \\
558: =\frac{i\Delta \exp\left[-i|n|z\
559: \mathrm{sgn}\left(\omega(q^2-\Delta^2)\right)\right]\mathrm{sgn}
560: \left(n(q^2-\Delta^2)\right)}
561: {\sqrt{(q^2-\Delta^2)(\omega^2-\Delta^2)}} \hat{\tau}_1\nn\\
562: +\frac{iq \exp\left[-i|n|z\
563: \mathrm{sgn}\left(\omega(q^2-
564: \Delta^2)\right)\right]\mathrm{sgn}(\omega)}{\sqrt{(q^2-\Delta^2)
565: (q^2-\omega^2)}} \hat{\tau}_3 \quad , \qquad \label{godd} \eea
566: 
567: where
568: 
569: \bea
570: z(k_y) &=& \tan^{-1}\sqrt{{q^2-\omega^2 \over \omega^2- \Delta^2 }}
571: \label{zomega} \enspace ,\\
572: \Delta(k_y) &=& 2 \Delta_0 \sin(k_y d) \label{dky} \enspace ,\\
573: q(k_y) &=& 4 t \cos(k_yd) \ , \label{qky} \eea and $\Delta(k_y)$
574: and $q(k_y)$ are the maximum gap $\Delta_\k$ and single-particle
575: spectrum $\xi_\k$ for fixed $k_y$ in 110 geometry. Note that here
576: the square root function takes positive values for positive
577: arguments, i.e. under the conditions
578: $|\Delta(k_y)|<|\omega|<|q(k_y)|$ or
579: $|q(k_y)|<|\omega|<|\Delta(k_y)|$. For $A^2-\omega^2<0$,\,the
580: branch $\sqrt{A^2-\omega^2}\to -i\ \mathrm{sgn}\omega\
581: \sqrt{\omega^2-A^2}$, for either $A=\Delta, q$.  In Fig.
582: \ref{fig:surf110} we show the LDOS spectra on the different
583: layers. Each layer is defined as an array of sites parallel to the
584: surface. Its index indicates its position; layer $n$ corresponds
585: to sites at a distance $nd$ away from the surface.
586: 
587: The Andreev bound states are manifested as zero-energy peaks in
588: the LDOS. For $\mu=0$, such  states are found in all odd layers,
589: and are absent on even ones, as seen in Figure \ref{fig:Andreev}.
590: This even-odd effect can be easily understood from the form of the
591: $T$-matrix and Green's functions Eqs. (\ref{t}-\ref{godd}).
592: \begin{figure}[tbh]
593: \begin{center}
594: \leavevmode
595: \includegraphics[clip=true,width=.8\columnwidth]{Andreev.eps}
596: \caption{Amplitude of Andreev state vs. distance $n=x/d$ from 110
597: surface for $\Delta_0=0.2t$, $\mu=0$ .} \label{fig:Andreev}
598: \end{center}
599: \end{figure}
600: For even $n$, the bare Green's function ${G}^{(0)}(n,k_y,\omega)$
601: is proportional to $\omega$ for low-frequencies. Then the
602: $T$-matrix $T \sim 1/\omega$ for small $\omega$.  Even though the
603: $T$-matrix has a pole at $\omega=0$, the product of two Green's
604: functions in the analog of Eq. \ref{greens}  decreases faster,
605: resulting in zero LDOS at zero frequency. As seen from the right
606: panel of Fig. \ref{fig:surf110}, the low-energy density of states
607: on even layers near the $(110)$ surface is substantially less than
608: in the bulk, where it varies linearly with sufficiently low
609: energy. Furthermore, the amplitude of the gap features in the LDOS
610: is noticeably suppressed with decreasing distance from the
611: surface. These features of the LDOS can be understood based on a
612: simple relation between the Green's function ${\hat G}(n'=n, k_y,
613: \omega)$ of the half-space and the bare Green's function ${\hat
614: G}^{(0)}(n-n=0, k_y, \omega)$, which takes place for even $n=2m$ :
615: \begin{eqnarray}
616: &&\hat{{G}}(n'=n, k_y, \omega)=\Bigl\{1-\qquad \qquad \qquad
617: \qquad \qquad
618: \qquad \nonumber \\
619: &&-\exp\left[-2i|n|z\
620: \mathrm{sgn}\left(\omega(q^2-\Delta^2)\right)\right]\Bigr\}
621: {\hat G}^{(0)}(0, k_y, \omega) . \label{g2m}
622: \end{eqnarray}
623: The factor in the braces in Eq.(\ref{g2m}) controls the deviation
624: from the bulk behavior and diminishes with decreasing even $|n|$.
625: At low energies only narrow regions of $k_y$ near the center and
626: the edge of the Brillouin zone contribute to the LDOS. As a
627: result, for even layers the density of states goes as $|\omega|^3$
628: at low energies, with the main contributions arising from $k_y$
629: near the edge of the Brillouin zone.
630: 
631: For odd layers, the $\tau_1$ and $\tau_3$ components of
632: $\hat{{G}}^{(0)}(n=2m+1,k_y,\omega)$ are the ones that are
633: non-zero, and they approach a constant value as $\omega$ goes to
634: zero. So the pole in the $T$-matrix generates the peak in the
635: LDOS, associated with the zero-energy Andreev surface states. For
636: Andreev states, $z$ is an imaginary quantity. The size of the peak
637: decreases as the distance to the surface increases due to the
638: $e^{-|n {\mathrm Im} z|}$ factor in Eq.(\ref{godd}). For large
639: $|n|$, small $k_y$ dominates the integration over $k_y$ in the
640: LDOS and we obtain the following zero-energy asymptotic behavior
641: of the LDOS:
642: $\rho(\omega)=\dfrac{2t}{\pi\Delta_0n^2}\delta(\omega)$. The size
643: of the zero-energy peak in the LDOS $\propto n^{-2}$.
644: 
645: 
646: \subsection{$(210)$ surface}
647: It is useful to study a case intermediate between the standard
648: $(100)$ and $(110)$ surfaces to see what qualitatively new features
649: arise. From the usual quasiclassical viewpoint, the weight of the
650: zero-energy Andreev $(210)$-surface states should be finite, but
651: smaller than for the $(110)$ surface because the phase space for
652: which the reflecting quasiparticle experiences a sign change of the
653: order parameter is reduced. The tight-binding model leads
654: to a more complicated dependence of the weight of the zero-energy
655: states on the surface orientations relative to the crystal axes.
656: Thus, for the $(210)$ surface the model shows at half-filling
657: no zero-energy Andreev states at all. We associate this discrepancy,
658: in particular, with the difference between reflection channels
659: incorporated in the two approaches.
660: 
661: Standard quasiclassical considerations imply that parallel to a
662: smooth surface the momentum component $k_y$ is conserved in a reflection
663: event, and only conventional specular reflection takes place. A
664: tight-binding model shows that, generally speaking, this is not the
665: case, since the {\it crystal momentum} component $k_y$ can also change in a
666: reflection process by a reciprocal crystal vector along the surface.
667: Due to a difference between reciprocal crystal vectors at the surface
668: and in the bulk, the momentum acquired by a quasiparticle in a
669: reflection event can be physically distinguished in the bulk from
670: that of specularly reflecting quasiparticle. Hence, specific crystal
671: periodicity along a particularly oriented surface can result in
672: additional channels for quasiparticle reflection, if there is a reflected
673: state $k_{x}(k_y),\, k_y$ on the Fermi surface corresponding to the
674: $k_y$ surface Umklapp process \cite{gantmakher}.
675: 
676: In this case the Fermi surface, considered as a part of the surface
677: adapted Brillouin zone, should exhibit multiple values of outgoing
678: $k_{F,x}$ for a fixed value of $k_{F,y}$. This situation is realized
679: for the comparatively complicated multisheet structure of the Fermi
680: surface in the surface-adapted Brillouin zone for the (210) surface,
681: as obtained in Fig.9 of Ref.\onlinecite{pairor02}. A strong dependence
682: of the particular shape of the surface-adapted Brillouin zone, as well
683: as the Fermi surface in the zone, on surface orientation is an important
684: characteristic feature of the tight-binding models \cite{Walker,pairor02}.
685: Within the quasiclassical theory, the shape of the Fermi surface is usually
686: considered as independent of surface orientations relative to the
687: crystal axes. We note, that the zero-energy surface states, as a rule,
688: do not exist within the quasiclassical approach,
689: if multiple channels for reflection of quasiparticles from an impenetrable
690: surface are assumed. In this subsection, we now study the LDOS for
691: quasiparticle spectra obtained with the tight-binding model at half-filling
692: for (210) surface.
693: 
694: Technically, we now need  to solve for the Green's function in the
695: presence of two impurity lines (Fig. \ref{fig:surf210}).
696: In the general case of two parallel impurity lines, we cut the
697: crystal at an orientation given by ($hl0$) and introduce an
698: impurity potential
699: \bea
700: U(\r) = U_0 \sum_j \delta(\r- \R_j) \enspace ,
701: \eea
702: where $\R_j$ are the points on the two impurity lines, at the
703: location of the boundaries. The first boundary is defined to be
704: located at $x=0$ with impurity sites $\R_j = j d^{-1} \hat{y}$
705: and the other is parallel and located at $x=(N+1)d$ with
706: $\R_{j} = (N+1)d \hat{x} + (c+jd^{-1})\hat{y}$, giving a total
707: of $N$ free chains. Here $c$ is a shift along $y$-axis of sites
708: on the $(N+1)$-th chain relative to the sites on the $0$-th chain.
709: For the special case of the $(210)$ surface, we need two
710: adjacent lines at $x=0$ and $x=-d$, so formally this corresponds
711: to the case of $N=-2$.
712: 
713: The equation for the $T$-matrix (Eq. \ref{tmat}) for this impurity
714: potential can be solved by choosing the ansatz
715: \bea
716: \hat T(k_x,k_x^{\prime}, k_y) &=& \hat t_0 + \hat t_1 e^{i (N+1)d k_x} +
717: \hat t_2 e^{-i(N+1)d k_x^{\prime}} \nonumber \\
718: &&+ \hat t_3 e^{i(N+1)d (k_x-k_x^{\prime})} \enspace .
719: \eea
720: In the limit of infinitely strong impurity potential this gives
721: \bea
722: \hat t_0 &=& {-\hat{G}^{(0)}(0)^{-1} \over 1 - \hat{ G}^{(0)}(0)^{-1} \hat{
723: G}^{(0)}(-N-1)\hat{ G}^{(0)}(0)^{-1} \hat{ G}^{(0)}(N+1)} \nonumber \\
724: \hat t_1 &=& {\hat{ G}^{(0)}(0)^{-1} \hat{ G}^{(0)}(N+1) \hat{
725: G}^{(0)}(0)^{-1} \over 1 - \hat{ G}^{(0)}(0)^{-1} \hat{
726: G}^{(0)}(-N-1)\hat{ G}^{(0)}(0)^{-1} \hat{ G}^{(0)}(N+1)}\nonumber \\
727: \hat t_3 &=& - {\hat{ G}^{(0)}(0)^{-1} \over  1 - \hat{
728: G}^{(0)}(0)^{-1} \hat{ G}^{(0)}(N+1) \hat{ G}^{(0)}(0)^{-1} \hat{
729: G}^{(0)}(-N-1)} \nonumber \\
730: \hat t_2 &=& {\hat{ G}^{(0)}(0)^{-1} \hat{ G}^{(0)}(-N-1) \hat{
731: G}^{(0)}(0)^{-1} \over 1 - \hat{ G}^{(0)}(0)^{-1} \hat{
732: G}^{(0)}(-N-1) \hat{ G}^{(0)}(0)^{-1} \hat{ G}^{(0)}(N+1)}
733: \nonumber
734: \eea
735: The Green's function in this case can then be
736: written as:
737: $$ \hat{ G}\!(n, n^{\prime}) = \hat{G}^{(0)}\!(n-n^{\prime}) -
738: \left(\hat{ G}^{(0)}\!(n) \ \ \ \hat{
739: G}^{(0)}\!(n-N-1) \right)\times$$
740: \bea
741: \!\!\left(\!\begin{array}{cc}\! \hat{ G}^{(0)}\!(0) & \hat{ G}^{(0)}\!(
742: \!- N\!-\!1)\\
743: \hat{ G}^{(0)}\!( N\!+\!1) & \hat{ G}^{(0)}\!(0)
744: \end{array} \!\right)^{\!-1}\!
745: \!\left(\!\begin{array}{c} \hat{ G}^{(0)}\!(-n^{\prime})\\
746: \hat{ G}^{(0)}\!(\! N\!+\!1\!-\! n^{\prime})
747: \end{array}\!\right)\enspace ,
748: \label{greenwire}
749: \eea
750: where we have not written the
751: dependence on $k_y$ and $\omega$ explicitly.
752: 
753: 
754: \begin{figure}[tbh!]
755: \begin{center}
756: \leavevmode
757: \includegraphics[width=.9\columnwidth]{surface210_S.eps}
758: \caption{Local density of states $\rho(x,\omega)$ versus
759: $\omega/\Delta_{max}$ for a closed(210) non-transparent surface at
760: half-filling model and $\Delta_0=0.2t$, $\mu=0$. Each curve
761: corresponds to a chain located at a distance $x/d$ to the
762: surface.} \label{fig:LDOS210}
763: \end{center}
764: \end{figure}
765: 
766: 
767: The equation for the bound state on (210) surface is
768: \bea
769: {\rm det} \left(\begin{array}{cc} \hat{{G}}^{(0)}(0) &
770: \hat{{G}}^{(0)}(1)\\
771: \hat{{G}}^{(0)}(-1) & \hat{{G}}^{(0)}(0) \end{array} \right) = 0
772: \enspace .
773: \label{det210}
774: \eea
775: 
776: 
777: Fig \ref{fig:LDOS210} displays the LDOS spectra calculated on
778: several layers for (210) surface.  There is no zero-energy peak
779: observed on any chain at all, for the reasons explained in the
780: beginning of this subsection. The peaks in the LDOS, seen close to
781: the (210) surface at energies $\pm 0.5\Delta_{max}$, originate
782: from the gap features taken for the momentum along the surface
783: normal: $\Delta(k_{F,x}, k_{F,y}=0)=0.5\Delta$. The peak at
784: $\omega=\Delta(k_{F,x},0)$ arises for the homogeneous model of
785: the order parameter, while for the self-consistent spatially
786: dependent order parameter it lies slightly below $\Delta(k_{F,x},0)$.
787: It is associated with the
788: surface Andreev states and decays in the depth of the
789: superconductor. These peaks have been theoretically found first in
790: Ref.\ \onlinecite{bbs97} with a continuous model and then also for
791: the conductance with a lattice model \cite{pairor02}. We notice
792: that the conductance spectrum shown in Fig.14 of
793: Ref.\onlinecite{pairor02}, is in agreement with the LDOS on the
794: first chain (at $x=d$) in Fig.\ref{fig:LDOS210}. The variations of
795: the LDOS from chain to chain, which accompany a large-scale
796: behavior, are the Friedel-like oscillations.
797: 
798: \section{Half-filled Wires}
799: 
800: We now consider wires where the second line of impurities confines
801: the system to a finite width, i.e. we restrict ourselves to the
802: cases with surface normal along the $(100)$ and $(110)$
803: directions. Semiclassically, a quasiparticle will go through
804: multiple scatterings, bouncing back and forth between the two
805: walls. Hence we expect that the interplay between Andreev
806: reflection, taking place due to sign change of the order
807: parameter, and the energy discretization, due to finite wire
808: width, to yield novel features. We again model the surfaces by
809: introducing impurities on the appropriate sites to completely
810: isolate the wires. The corresponding Green's function is given by
811: Eqs.(\ref{greenwire}). The bound state energies are determined by
812: the equation \bea {\rm det} \left(\begin{array}{ll}
813: \hat{{G}}^{(0)}(0) &
814: \hat{{G}}^{(0)}(-N-1)\\
815: \hat{{G}}^{(0)}(N+1) & \hat{{G}}^{(0)}(0) \end{array} \right) = 0
816: \enspace .
817: \label{det}
818: \eea
819: 
820: 
821: \subsection{(100) wires}
822: 
823: 
824: Ziegler et al.\cite{zieglerwire}  investigated the problem of
825: (100) $d$-wave quantum wires in some detail, and discovered the
826: existence of a number-parity effect as a function of the width $N$
827: of a mesoscopic $d$-wave wire: a finite total DOS is found at the
828: Fermi level for odd $N$ and zero DOS (with a full gap in the
829: excitation spectrum, not a $d$-wave like gap) is found for even
830: $N$, at least for a simple tight-binding band at half-filling. The
831: differences between even and odd $N$ were shown to survive for
832: more general bands, as well.   For completeness, we reproduce,
833: using our approach, some of their LDOS spectra in Fig.
834: \ref{fig:wires100}.
835: 
836: \begin{figure}[tbh]
837: \begin{center}
838: \leavevmode
839: \includegraphics[clip=true,width=.7\columnwidth,angle=90]{strips_100_S.eps}
840: \caption{Local density of states $\rho(x,\omega)$ vs.
841: $\omega/\Delta_{max}$ for a $(100)$ wire of width $N=4$ (left panels)
842: and $N=5$ (right panels), using $\Delta_0=0.2t$, $\mu=0$. }
843: \label{fig:wires100}
844: \end{center}
845: \end{figure}
846: 
847: 
848: \subsection{(110) wires}
849: 
850: \subsubsection{Normal metal wires}
851: \label{sec:nwires} Based on intuition from the surface case, we
852: would expect that Andreev states play an important role  in (110)
853: wires, with geometry shown in Figure \ref{fig:strip110}. A crucial
854: question which arises in the following discussion is, how does one
855: identify a subgap state of true Andreev character? By merely
856: measuring the LDOS with an STM, for example, one may see several
857: peak structures, not all of which will be related to Andreev
858: reflections at the surfaces. One set of candidate states which
859: needs to be investigated first is the set of discrete dispersive
860: (with respect to $k_y$) levels which arise simply because of the
861: finite wire width.  These are of course present already in the
862: normal state wire.
863: 
864: The $k_x$-integrated bare normal state Green's function for an
865: infinite lattice above $T_c$ takes the form : \bea
866: {G}^{(0)}_{11}(n,k_y,\omega)&=&\frac{d}{2\pi}\int^{\pi/d}_{-
867: \pi/d}\frac{e^{ik_xnd}dk_x} {(\omega+i0)-\xi_\k}\nonumber\\
868: &=& - {i\exp [i|n| \arccos (-\omega/q(k_y))]\over
869: \sqrt{q^2(k_y)-\omega^2}}\, ,\eea  where $ q(k_y)$ is defined in
870: Eq.(\ref{qky}). The full Green's function for a (110) wire may
871: then be obtained by solving the $T$-matrix equation as
872: \be
873: G_{11}(n,n';k_y,\omega)=\frac{2\sin[n_{min}z]\sin[(n_{max}-N-
874: 1)z]}
875: {\sqrt{q^2(k_y)-\omega^2}\sin[(N+1)z]}
876: \label{gnm}
877: \ee
878: where $n,n'=1,2,...,N$, $z\equiv z(k_y,\omega)=\cos^{-1}(-\omega/q(k_y))$,
879: $n_{min}={\rm min}(n,n')$,  and $n_{max}={\rm max}(n,n')$.
880: 
881: \begin{figure}[tbh]
882: \begin{center}
883: \leavevmode
884: \includegraphics[width=.7\columnwidth]{strip110.eps}
885: \caption{Geometry of 110 wire (sites are filled circles) of width
886: $N$ bounded by 2 impurity lines (filled squares).  Wire is
887: infinite in both $y$ and $-y$ directions.} \label{fig:strip110}
888: \end{center}
889: \end{figure}
890: 
891: 
892: For every value of $k_y$, there generally exists a series of
893: eigenvalues which are the poles of the Green's function. They may
894: be obtained by simply solving
895: \begin{equation}
896: \sin[(N+1)z(k_y,\omega)]=0 \enspace ,
897: \label{eigennormal}
898: \end{equation}
899: which yields:
900: \begin{equation}
901: \omega_\nu(k_y)=-q(k_y)\cos\frac{\pi \nu}{N+1},~~
902: \nu=0,1,...,N+1 \ .
903: \label{modefreqnormal}
904: \end{equation}
905: This gives $N$ branches of solutions for $\nu = 1, ..., N$,
906: distributed symmetrically with respect to the Fermi level, and two
907: special solutions ($\nu=0$ and $N+1$) with $\omega=\pm q(k_y)$
908: which we refer to as defining the "effective band edge". One
909: interesting feature is the existence in the normal metal wire of
910: dispersionless zero-energy  quasiparticle states (ZES) which form
911: the branch $\nu=(N+1)/2$ when the width $N$ of the wire is odd.
912: Since the group velocity vanishes for dispersionless states, they
913: are always localized. We note that the case $N=1$ has only the
914: obvious dispersionless ZES since no transport is allowed with only
915: nearest neighbor hopping. In the general $N$=odd case, it appears
916: that there is always exactly one such localized state (doubly
917: degenerate in  particle-hole space) for any fixed $k_y$. A
918: deviation from half filling shifts the zero-energy states to
919: $-\mu$. Thus, for positive $\mu$ they are the hole states, while for
920: $\mu<0$ - electron states.
921: 
922: \begin{figure}[tbh]
923: \begin{center}
924: \leavevmode
925: \includegraphics[clip=true,width=.9\columnwidth]{strip_110_5_NS.eps}
926: \caption{Local density of states  for a (110)  wire of width
927: $N=5$, with $\Delta_0=0.2t$, $\mu=0$. Left: normal state,
928: $\rho(x,\omega)$ vs. $\omega/\Delta_{max}$; right: superconducting
929: state, $\rho(x,\omega)$ vs. $\omega/\Delta_{max}$.}
930: \label{fig:N5wires}
931: \end{center}
932: \end{figure}
933: 
934: The contribution of each of these states to the LDOS can be
935: estimated by examining the residue near the pole, where the
936: Green's function can be approximated as:
937: \begin{equation}
938: G(n,n^{\prime};k_y,\omega)
939: \approx\frac{Q_\nu(n,n^{\prime})}{\omega-\omega_\nu(k_y)},
940: \end{equation}
941: with the residue given by:
942: \begin{equation}
943: Q_\nu(n,n')=\frac{2\sin\frac{n\pi \nu}{N+1}\sin\frac{n^{\prime}\pi
944: \nu}{N+1}}{N+1},\label{weightn0}
945: \end{equation}
946: where $n$ is the index of the layers. Note from
947: (\ref{weightn0}) and the spectral representation of the Green's
948: function near a pole, it is easy to read off the quasiparticle
949: wavefunctions as $\psi_\nu(n) =\sqrt{2/(N+1)}\sin[n\pi\nu/(N+1)]$,
950: i.e. just the wave functions of a free particle confined to a box
951: of width $(N+1)d$.
952: 
953: \begin{figure}[tbh]
954: \begin{center}
955: \leavevmode
956: \includegraphics[clip,width=.9\columnwidth]{strip_110_4_NS.eps}
957: \caption{LDOS for a (110)  wire of width $N=4$ with
958: $\Delta_0=0.2t$, $\mu=0$. Left: normal state, $\rho(x,\omega)$ vs.
959: $\omega/\Delta_{max}$; right: superconducting state,
960: $\rho(x,\omega)$ vs. $\omega/\Delta_{max}$.} \label{fig:N4wires}
961: \end{center}
962: \end{figure}
963: 
964: It is easy to check that for any $n$, $n'$ the residue $Q_\nu$
965: vanishes everywhere for the "states" at the effective bandwidth $\nu=0$ or
966: $N+1$; we therefore do not discuss them further, but focus on the
967: $N$ branches with finite residue. For these states the
968: residue can vanish locally on chains with number $n$, for
969: which the quantity $n\nu/(N+1)$ is an integer. The odd $N$ ZES,
970: for example, has a residue
971: \begin{equation}
972: Q_{\frac{(N+1)}{2}}(n,n)=\frac{2(\sin\frac{n\pi }{2})^2}{N+1}
973: \enspace .
974: \label{weightnn}
975: \end{equation}
976: As is seen from Eq.(\ref{weightnn}), the probability density of
977: the ZES oscillates with a period $2d$, taking finite values only
978: for odd $n$ and vanishing on all nearest neighbor sites (~where $n$
979: is even~). This is also valid for the states with energies $\pm\mu$
980: in the case of the deviation from half-filling and ensures no transport
981: with nearest neighbor hopping for quasiparticles with energies $\pm\mu$.
982: 
983: All these states are not surface bound states, as one can easily
984: check that their amplitude does not decay across the wire. The
985: quasiparticle spectrum is discrete (for fixed $k_y$) due to the
986: finite width of the wire and transforms into conventional
987: continuous quasiparticle spectrum in the massive normal metal in
988: the limit $N\to\infty$. The momentum resolved LDOS in the normal
989: metal wire with discrete dispersive states takes the form
990: \begin{equation}
991: \rho(n,k_y,\omega)=\sum\limits_{\nu=1}^{N}Q_{\nu}(n=n')\delta(\omega-
992: \omega_{\nu}(k_y)) \enspace .
993: \end{equation}
994: The integration over $k_y$ gives the LDOS:
995: \begin{equation}
996: \rho(n,\omega)=\frac{1}{2\pi d}\sum\limits_{\omega_{\nu}(k_y)=\omega}
997: \dfrac{Q_{\nu}(n,n)}{\left|\dfrac{d\omega_{\nu}}{dk_y}\right|} \enspace ,
998: \label{rhon}
999: \end{equation}
1000: where the sum is taken over those $k_y$ and $\nu$, which satisfy
1001: the equation $\omega_{\nu}(k_y)=\omega$.
1002: 
1003: The position of peaks in LDOS are then determined by extrema of
1004: the dispersive branches (\ref{modefreqnormal}), taking place at
1005: $k_y=0$: \be \omega_{\nu,peaks}=-4t\cos\frac{\pi \nu}{N+1},~~
1006: \nu=1,...,N \ . \label{energypeaks}
1007: \end{equation}
1008: 
1009: The peaks corresponding to the $N$ normal metal wire states are
1010: seen clearly in Figs. \ref{fig:N5wires} and \ref{fig:N4wires} at
1011: the eigenfrequencies given by Eq. (\ref{energypeaks}). It is easy
1012: to check that the weights agree with Eq. (\ref{weightn0}). The
1013: LDOS for the normal metal wires qualitatively differs from the
1014: LDOS for bulk normal metals with the nearest neighbor hopping on
1015: the square. The zero-energy peak (the Van Hove singularity) in the
1016: bulk metal is symmetric as a function of the energy and its log
1017: singularity is much broader than the  $\delta$-like peaks we find
1018: here for wires.
1019: 
1020: \subsubsection{Superconducting wires}
1021: \label{sec:swires}
1022: 
1023: The $T$-matrix and Green's function equations are necessarily more
1024: complicated in the presence of superconductivity, but they are
1025: still tractable in the $(110)$ case. The bare Green's functions
1026: $\hat{G}^{(0)}$ are given by Eqs. (\ref{geven}) and (\ref{godd}),
1027: as before. One must then solve Eq. (\ref{det}), which applies to
1028: any situation which requires two lines of impurities, for the
1029: eigenenergies $\omega$. Similar to the normal metal case, there are
1030: special solutions of Eq. (\ref{det}): $\omega=\pm\Delta(k_y)$,
1031: $\omega=\pm q(k_y)$ for any $N$ and $\omega=0$ for even $N$. They
1032: are  poles of the $T$-matrix, but do not correspond to the poles
1033: of the full Green's function.
1034: 
1035: The behavior of subgap surface states on the narrow
1036: superconducting wire differs qualitatively from the case of the
1037: superconducting halfspace due to the interference of the wave
1038: functions of the states on both surfaces. Since the zero-energy
1039: peak in the LDOS for the half-filled surface vanishes on each even
1040: chain (~see Figs.(\ref{fig:surf110}), (\ref{fig:Andreev})), the
1041: spectrum of  Andreev states on the $(110)$ wires becomes strongly
1042: dependent on the parity of the number $N$ of chains in the
1043: half-filled wire. This effect is quite pronounced for wires whose
1044: width is the order of or less than the superconducting coherence
1045: length. As we demonstrate below, some new qualitative features
1046: arising in the superconducting state of the $(110)$ wires in the
1047: quasiparticle spectrum above $\Delta(k_y)$ (~see Eq. (\ref{dky}))
1048: can also be strongly dependent on the parity of $N$.
1049: 
1050: \vskip .2cm
1051: 
1052: {\it Odd $N$.} For a wire with odd $N$, only the Green's functions
1053: (\ref{geven}) with even arguments  $n=0, \pm(N+1)$ enter
1054: Eq.(\ref{det}).  Since at small frequencies these Green's function
1055: are $\propto \omega$, it is straightforward to show that the only
1056: subgap state is a dispersionless ZES $\omega=0$. Dispersive modes
1057: exist as well and take the form
1058: 
1059: \begin{eqnarray} \omega_{\nu}^2(k_y)=q^2(k_y)\cos^2\frac{\pi
1060: {\nu}}{N+1}+\Delta^2(k_y)\sin^2\frac{\pi
1061: {\nu}}{N+1}~, \nonumber\\
1062: ~~{\nu}=1,...,\frac{N-1}{2} \ .
1063: \label{dm}
1064: \end{eqnarray}
1065: 
1066: This exactly coincides with the quasiparticle spectrum in a bulk
1067: two-dimensional superconductor $\omega^2(k_y, k_x)=\xi^2(k_y, k_x)
1068: +\Delta^2(k_y,k_x)$, if the discrete values of momentum component
1069: across the wire $k_{x,\nu}=\pi{\nu}/(N+1)d$ are introduced. We
1070: note that the dispersionless zero-energy Andreev states are the
1071: only true subgap states in the spectrum, since the energy
1072: $|\omega_{\nu}(k_y)|$ of the dispersive states (\ref{dm}) lie
1073: above the respective value $|\Delta(k_y,k_x)|$ of the order
1074: parameter, for any $\nu$ and $k_y$.
1075: 
1076: \begin{figure}[htb]
1077: \begin{center}
1078: \leavevmode
1079: \includegraphics[width=.9\columnwidth]{N11_combi.eps}
1080: \caption{ Quasiparticle spectra and the LDOS for superconducting
1081: half-filled (110) wire with $N=11$.\ (a) Dispersive modes ($k_yd$
1082: vs. $\omega/t$) for $N=11$, $\Delta_0=0.2t$, $\mu=0$. (b) The
1083: blowup of the dispersive spectra near the edge of the Brillouin
1084: zone.\ (c) The LDOS for energies less than $\Delta_{max}$: n=2
1085: (solid line), n=1 (dashed line), where a broadening $\delta=.005t$
1086: has been used. }
1087: \label{figN11}
1088: \end{center}
1089: \end{figure}
1090: 
1091: There are, as in the normal state case, $N-1$ dispersive modes in
1092: addition to the ZES. They are doubly degenerate due to the
1093: particle-hole symmetry. This simple result can be understood as
1094: follows. For a fixed $k_y$ the problem reduces to a
1095: one-dimensional problem. The corresponding two-component
1096: Bogoliubov-DeGennes wave function takes its values on $N$ sites,
1097: resulting in $2N$ degrees of freedom in the system. With this
1098: point of view it appears natural that for fixed $k_y$ the total
1099: number of levels, which are twice degenerate, is $N$. The set of
1100: levels with positive energies for $N=11$ is represented in panel
1101: (a) of Fig.\ref{figN11}.
1102: 
1103: The LDOS for the superconducting wires with N=5 is shown in the
1104: right panel of Fig.~\ref{fig:N5wires}. The dispersionless ZES
1105: results in a pronounced peak at zero energy on odd layers.
1106: Furthermore, each extremum of the dispersive mode $\omega(k_y)$
1107: results in the peak in the LDOS at the energy
1108: $\omega_{k_{y,extr}}$. One series of  peaks is associated with the
1109: extrema at $k_y=0$. Since $\Delta(k_y=0)=0$, the peaks lie at the
1110: same positions as in the normal metal wire (\ref{energypeaks}).
1111: Although they are irrelevant to the superconducting properties of
1112: the wire, some of these peaks can lie at finite energies below the
1113: maximum of the gap function $\Delta_{max}=2\Delta_0$. For
1114: instance, the lowest position at finite energies of the peaks of
1115: this series is $4t\sin\frac{\pi}{N+1}$. This can be both above or
1116: below $\Delta_{max}$, depending on the ratio $\Delta_{max}/t$ and
1117: the wire width $N$. In contrast with the normal metal wires, the
1118: dispersive quasiparticle modes Eq.(\ref{dm}) in the
1119: superconducting wires have extrema also at the edge of the
1120: surface-adapted Brillouin zone $k_y=\pm \pi/(2d)$ \cite{ebz}. They
1121: are shown in panel (b) of Fig.\ref{figN11} for the wire with
1122: $N=11$. This leads to additional series of $(N-1)$ quasiparticle
1123: peaks in the LDOS for the superconducting wires, which lie below
1124: $\Delta_{max}$:
1125: 
1126: \begin{equation} \omega_{peaks}=\pm \Delta_{max}\sin\frac{\pi \nu}{N+1},~~
1127: \nu=1,...,\frac{N-1}{2} \ .
1128: \label{energypeaksodd}
1129: \end{equation}
1130: Panel (c) of Fig.\ref{figN11} displays the series of peaks in the LDOS for
1131: $N=11$.
1132: 
1133: The  dispersive states forming the nonzero low-energy peaks
1134: (\ref{energypeaksodd}) in LDOS, are {\it not} Andreev states. They lie
1135: below $\Delta(k_y)$ for $k_y$ near the edges of Brillouin zone, but they
1136: are situated above the bulk gap function $\Delta(k_x(\nu),k_y)$. The wave
1137: function for any of these states possesses the finite current of
1138: the probability density, while for Andreev states the total
1139: probability current vanishes.
1140: 
1141: While positions of the peaks are determined by the
1142: extrema of the dispersive energies, their weights in the LDOS
1143: are controlled also by quasiparticle wave functions, which form the
1144: quantity $Q(n,n)$ in Eq.(\ref{rhon}).
1145: For this reason the weights of the peaks can substantially
1146: differ on different layers and may vanish on some of them, quite
1147: analogously to the normal metal wires (see Eq.(\ref{weightn0})).
1148: Due to an interference from two surfaces, the wave function, taken
1149: on even layers, turns out to coincide with the respective wave
1150: function in the normal metal case (apart from a normalization
1151: constant). On the other hand, the wave function on odd layers is a
1152: superposition of electron-like and hole-like Bogoliubov
1153: quasiparticles on the wire.
1154: 
1155: \vskip .5cm {\it Even $N$.}
1156: In the case of even $N$ half-filled wires, the quasiparticle
1157: spectra become more complicated. The Green's functions
1158: (\ref{geven}), (\ref{godd}) with even ($n=0$) and odd
1159: ($n=\pm(N+1)$) arguments enter Eq.(\ref{det}), which can be
1160: reduced to  the following form
1161: \be q(k_y)\tan[(N+1) z]=
1162: \alpha\Delta(k_y)\tan z \enspace ,
1163: \label{z}
1164: \ee
1165: where $\alpha=\pm
1166: 1$, $0\le z\le\pi/2$. Solutions $z_{\nu,\alpha}(k_y)$ of
1167: Eq.\eqref{z} are directly associated with $\omega$, in accordance
1168: with Eq.\eqref{zomega}:
1169: \begin{equation}
1170: \omega^2_{\nu,\alpha}(k_y)=q^2(k_y)\cos^2z_{\nu,\alpha}(k_y)+
1171: \Delta^2(k_y)\sin^2z_{\nu,\alpha}(k_y) \enspace .
1172: \label{evenspectr}
1173: \end{equation}
1174: Here ${\nu}=1,...,\frac{N}{2}$ and $\alpha$ are the indices of the
1175: solution. Comparing Eqs.\eqref{evenspectr}, \eqref{dm} for the
1176: spectra of odd and even wires, one can see that $z/d$ plays the
1177: role of effective discrete values of the momentum component $k_x$
1178: (at fixed $k_y$) across the wire. Eq.\eqref{z} can be transformed
1179: to a polynomial equation in $\tan z$ of the $N$-th degree, if one
1180: excludes the special solutions $\omega=\pm\Delta(k_y), \pm q(k_y)$
1181: mentioned above. Hence, for a fixed $\alpha$ there are exactly
1182: $N/2$ positive and $N/2$ negative solutions for $\omega$,
1183: describing $N$ dispersive branches of the quasiparticle spectrum.
1184: Explicit analytical form of the spectra can be easily found from
1185: Eqs.\eqref{evenspectr}, \eqref{z} in the particular cases $N=2$
1186: \be
1187: \omega_{\alpha}=\pm\dfrac{1}{2}\Bigl(q(k_y)+\alpha\Delta(k_y)\Bigr),
1188: \label{2} \ee and $N=4$ ($\nu=\pm$, $\alpha=\pm$)
1189: \begin{eqnarray}
1190: &\omega^2_{\nu,\alpha}(k_y)=\Delta^2(k_y)+\dfrac{1}{8}q(k_y)\Bigl(q(k_y)+
1191: \alpha\Delta(k_y)\Bigr)\Biggl[3- \nonumber \\
1192: & -5\alpha\dfrac{\Delta(k_y)}{q(k_y)}
1193: +\nu\,\sqrt{5-6\alpha\dfrac{\Delta(k_y)}{q(k_y)}+
1194: 5\dfrac{\Delta^2(k_y)}{q^2(k_y)}}\,\Biggr] . \label{4}
1195: \end{eqnarray}
1196: Eq.\eqref{4} is defined for all $k_y$ in the Brillouin zone for
1197: which $\omega^2_{\nu,\alpha}(k_y)$ is positive. Eqs.\eqref{2},
1198: \eqref{4} describe the quasiparticle spectra for
1199: non-self-consistent wires with small numbers of chains. As we will
1200: show in Sec. \ref{sec:self}, the self-consistent treatment of the
1201: problem can lead to important modifications of the results, at
1202: least if the width of the wire is less than or of  order  the
1203: superconducting coherence length.
1204: 
1205: \begin{figure}[tbh]
1206: \begin{center}
1207: \leavevmode
1208: \includegraphics[width=.9\columnwidth]{N10_combi.eps}
1209: \caption{Quasiparticle spectra and the LDOS for superconducting
1210: half-filled (110) wire with $N=10$.\ (a) Dispersive modes ($k_yd$
1211: vs. $\omega/t$) for $N=10$, $\Delta_0=0.2t$, $\mu=0$.\ (b) The
1212: blowup of the spectra  near the edge of the Brillouin zone.\ (c)
1213: The LDOS for energies less or equal to $\Delta_{max}$ : n=2 (solid
1214: line), n=1 (dashed line), where a broadening $\delta=.005t$
1215: has been used. }
1216: \label{figN10}
1217: \end{center}
1218: \end{figure}
1219: 
1220: Whereas in the normal metal state, $\nu$ is the only index of the
1221: solution (for a given $k_y$), in the superconducting even-$N$ wire
1222: each branch with fixed $\nu$ splits into two, corresponding to two
1223: values of the index $\alpha$, giving a  total  of $2N$ distinct
1224: branches. Since the order parameter vanishes for $k_y=0$, the
1225: splitting is absent in this particular case. The splitting is
1226: associated with the symmetry breaking of the eigenstates with
1227: respect to the sign reversal of $k_y$. In the normal  even-$N$
1228: wires, as well as in the odd $N$ wires (both in the normal metal
1229: and in the superconducting states), each separate branch of the
1230: spectra $\omega_{\nu}(k_y)$ is an even function of $k_y$. This is
1231: not the case, however, for the even $N$ superconducting $(110)$
1232: wires. As directly seen from Eq.\eqref{z} and the relations
1233: $q(-k_y)=q(k_y)$, $\Delta(-k_y)=-\Delta(k_y)$, the solutions of
1234: Eq.\eqref{z} $z_{\nu,\alpha}(k_y)$ with fixed $\alpha$ are not odd
1235: or even functions of $k_y$, due to the nonzero right hand side of
1236: Eq.\eqref{z} in the superconducting state. Evidently, each
1237: separate branch of the spectra $\omega_{\nu,\alpha}(k_y)$ possesses
1238: the same property. The whole spectrum, however, remains symmetric
1239: with respect to $k_y\to-k_y$, since the branches with positive and
1240: negative $\alpha$ simply interchange with each other under this
1241: transformation. Spectra of the half-filled $N=10$ superconducting wire
1242: with positive energies are shown in panel (a) of
1243: Fig.~\ref{figN10}. They agree with the above discussion. For
1244: branches with higher energies, the splitting is less than for lower
1245: branches. For small splittings, $\delta\omega\propto \Delta/(N+1)$.
1246: 
1247: The splitting manifests itself by slightly shifting the positions
1248: of peaks in the LDOS, due to respective shifts of extrema of
1249: dispersive quasiparticle energies. In contrast with the double
1250: number of  extrema, the total number of the peaks in the LDOS does
1251: not change. Any peak in the LDOS originates from the two extrema,
1252: situated symmetrically with respect to the sign of $k_y$. Each
1253: pair of extrema belongs to two spectral branches, which differ
1254: only by their index $\alpha=\pm1$. If we disregard the splitting
1255: of the quasiparticle spectra, the positions of the peaks in the
1256: LDOS coming, for example, from the edge of the Brillouin zone, are
1257: as follows
1258: \begin{equation} \omega_{peaks}=\pm \Delta_{max}\cos\frac{\pi \nu}{N+1},~~
1259: \nu=1,...,\frac{N}{2} \ .
1260: \label{energypeakseven}
1261: \end{equation}
1262: The splitting shifts the peak positions towards slightly lower
1263: energies $|\omega|$, as compared with those in Eq.\eqref{energypeakseven}.
1264: The respective extrema of dispersive energies are shifted towards lower
1265: values of $|k_y|$ from $k_y=\pm\pi/(2d)$. These low-energy peaks in the
1266: LDOS are shown in panel (c) of Fig.~\ref{figN10}.
1267: 
1268: The other peaks in the LDOS come, neglecting the splitting, from
1269: the center of the Brillouin zone. Since the order parameter
1270: vanishes in the center of the zone, the positions of the peaks are
1271: still the same as in Eq.\eqref{energypeaks} for the normal metal
1272: state. The splitting, taking place in the superconducting state,
1273: slightly shifts the extrema of the spectra from the center of the
1274: zone, so that the peaks move to a larger energy values $|\omega|$,
1275: as compared with those in Eq.\eqref{energypeaks}.
1276: 
1277: The LDOS for the superconducting wire with $N=4$ is shown in the
1278: right panel of Fig.\ref{fig:N4wires}. In addition to the peaks at
1279: finite energies, discussed above, there is also a well pronounced
1280: zero-energy peak there (see also panel (c) of Fig.~\ref{figN10}).
1281: This peak originates from the extrema of two dispersive branches
1282: of Andreev states, which take place at $\pm k_{y,0}d=\pm
1283: \tan^{-1}(2t/\Delta_0)$, where the relation
1284: $q(k_{y,0})=\Delta(k_{y,0})$ holds. Indeed, as  follows from
1285: Eqs.(\ref{z}), (\ref{zomega}) and, in particular, directly seen
1286: from Eq.(\ref{4}), the energy and its derivative equal zero for
1287: $k_y=\pm k_{y,0}$. It turns out that the multiplicity of zeroes
1288: of the lowest dispersive branches of states at the extrema $k_y=\pm k_{y,0}$ is $N/2$,
1289: i.e. $\omega\propto |k_y-k_{y,0}|^{N/2}$ in the close vicinity of $k_{y,0}$.
1290: This follows directly, for instance, from Eq.(\ref{omegaa}) given below.
1291: We notice, that the lowest dispersive curve in panel (b) of Fig.~\ref{figN10}
1292: manifests very slow change of energy when it touches the zero-energy line.
1293: This agrees with our expectation $\omega\propto |k_y-k_{y,0}|^{5}$
1294: for the (110) wire with $N=10$. The respective peak in the LDOS of wires with
1295: large even $N$ diverges at $\omega=0$, but is not a delta-like peak as it is
1296: for wires with odd $N$.
1297: 
1298: The identification of the Andreev nature of the quasiparticle bound
1299: states in confined geometries turns out to be a nontrivial
1300: problem, at least in the case of half-filled (110) wires with even
1301: $N$, where there is a symmetry breaking with respect to
1302: $k_y\to-k_y$. As seen from Eqs.(\ref{z})-(\ref{4}), the asymmetry
1303: can be associated with the order parameter behavior
1304: $\Delta(-k_y)=-\Delta(k_y)$ and with the sensitivity of the
1305: quasiparticle energies to the $\pi$-shift of the order parameter
1306: phase. A dependence of the dispersive energy curve on the order parameter
1307: phase is usually an intrinsic feature of Andreev bound states only.
1308: We consider the vanishing of  the probability density current as an
1309: defining property of  Andreev bound states. These states should
1310: be able to carry finite electric current, however. In odd $N$ wires
1311: only zero-energy dispersionless quasiparticle states satisfy the
1312: above requirements. In even $N$ $(110)$ wires the zero-energy
1313: dispersionless states do not arise at all. The interference of
1314: wave functions located near two surfaces of the even $N$ wire
1315: induces dispersive Andreev branches $\pm\omega_A(k_y)$, which
1316: transform into the zero-energy states in the limit of infinitely
1317: large $N$. Moreover, all quasiparticle states in the half-filled
1318: $(110)$ superconducting wire with even $N$, even those lying above
1319: the gap, turn out to satisfy the above mentioned conditions, i.e.
1320: possess the properties of Andreev bound states. As an example, we
1321: describe below in detail the structure of the lowest dispersive
1322: quasiparticle branch, which we denote as Andreev branch
1323: $\omega_A(k_y)$, although its energy varies with $k_y$  over a
1324: large range, including both the subgap and the supergap regions
1325: (see, for example, Fig.~\ref{figN10}).
1326: 
1327: For those $k_y$ where ${\rm
1328: min}\{|\Delta(k_y)|,q(k_y)\}<|\omega_A(k_y)| <{\rm
1329: max}\{|\Delta(k_y)|,q(k_y)\}$, the wave function of the state can
1330: be written on odd layers ($n=2m+1$) as: \be \left(\!
1331: \begin{array}{c} u\\v \end{array}\!\right)= C{\rm sgn}\omega_A
1332: \sin[(\!N\!+\!1\!-\!n\!)z] {1\choose {i{\rm sgn}k_y}}
1333: \label{aodd}
1334: \ee
1335: and on even layers $n=2m$
1336: \be
1337: \left(
1338: \begin{array}{c} u\\v\end{array}\right)= C(-1)^{N/2}\sin(nz)
1339: {1\choose {-i {\rm sgn}k_y}} \enspace .
1340: \label{aeven}
1341: \ee
1342: 
1343: In the range of $k_y$ for which
1344: $|\omega_A(k_y)|<{\rm min}\{|\Delta(k_y)|, q(k_y)\}$,
1345: the quantity $z$, entering Eq.\eqref{evenspectr}, becomes imaginary.
1346: Under the condition $|\omega_A(k_y)|<|\Delta(k_y)|<q(k_y)$
1347: the wave function takes the following form on odd layers
1348: ($n=2m+1$):
1349: \be
1350: \left(\! \begin{array}{c} u\\v \end{array}\!\right)=
1351: C_1 (-1)^m {\rm sgn}\omega_A
1352: \sinh[(\!N\!+\!1\!-\!n\!)z_1] {1\choose {i{\rm sgn}k_y}}
1353: \label{aodd1}
1354: \ee
1355: and on even layers $n=2m$
1356: \be
1357: \left( \begin{array}{c} u\\v\end{array}\right)=
1358: C_1(-1)^m\sinh(nz_1) {1\choose {-i {\rm sgn}k_y}} \enspace .
1359: \label{aeven1}
1360: \ee
1361: Analogously, if $|\omega_A(k_y)|<q(k_y)<|\Delta(k_y)|$, the wave function
1362: on odd layers ($n=2m+1$)
1363: \be
1364: \left(\! \begin{array}{c} u\\v \end{array}\!\right)=
1365: C_2 {\rm sgn}\omega_A
1366: \sinh[(\!N\!+\!1\!-\!n\!)z_1] {1\choose {i{\rm sgn}k_y}}
1367: \label{aodd2}
1368: \ee
1369: and on even layers $n=2m$
1370: \be
1371: \left( \begin{array}{c} u\\v\end{array}\right)=
1372: C_2 \sinh(nz_1) {1\choose {-i {\rm sgn}k_y}} \enspace .
1373: \label{aeven2}
1374: \ee
1375: Here $C$, $C_1$ and $C_2$ are normalization constants,
1376: $z_1=|{\rm Im}z(\omega_A(k_y),k_y)|$, $z$ is defined in
1377: Eq.\eqref{zomega} and taken at $\omega=\omega_A(k_y)$.
1378: 
1379: The condition $|u(n,k_y)|=|v(n,k_y)|$, which is valid for all
1380: solutions Eqs.~\eqref{aodd} -- \eqref{aeven2},  results in
1381: zero total probability current density, while the electric current
1382: does not vanish for the given branch. This ensures the Andreev
1383: character, as defined above, of all the states in even $N$ wires,
1384: regardless of whether their energies lie above or below the gap.
1385: 
1386: The quasiparticle states which belong to the same dispersive
1387: branch $\omega_A(k_y)$ can have their wavefunctions both
1388: symmetrically decaying in the depth of the wire
1389: (Eqs.~\eqref{aodd1} -- \eqref{aeven2}) or oscillating and forming
1390: standing waves across the wire (Eqs.~\eqref{aodd}, \eqref{aeven}).
1391: In particular, the amplitude of the wave functions in
1392: Eqs.~\eqref{aodd1} -- \eqref{aeven2} takes its maximum value on
1393: layers $n=1$ and $n= N$, and manifests  Friedel-like oscillations,
1394: as the layer index changes from odd to a neighbor even value or
1395: vice versa. On a larger scale the amplitude decays in the bulk of
1396: the wire symmetrically with respect to two surfaces.
1397: 
1398: It is instructive to follow how the above results transform into
1399: the well known dispersionless zero-energy Andreev surface states
1400: in the limit of large $N$. For sufficiently large width of the
1401: wire compared with the coherence length, and for those $k_y$ where
1402: the wavefunction decays in the depth of the wire, the dispersive
1403: energy $\omega_A(k_y)$ takes a relatively simple form:
1404: \begin{eqnarray}
1405: \omega_A(k_y)=\pm\dfrac{2\Delta(k_y)
1406: q(k_y)}{\sqrt{|q^2(k_y)-\Delta^2(k_y)|}}\times
1407: \qquad \qquad \qquad \quad \nonumber\\
1408: \times\exp\left[-(N+1){\sinh}^{-1}\dfrac{{\rm min}\{|\Delta(k_y)|,
1409: q(k_y)\}}{\sqrt{|q^2(k_y)-\Delta^2(k_y)|}}\right] . \label{omegaa}
1410: \end{eqnarray}
1411: It follows from Eq.\eqref{omegaa}, that the energy of the Andreev
1412: states is exponentially small and vanishes in the limit of
1413: infinitely large $N$. With decreasing energy, the range of $k_y$
1414: where the wave function oscillates (see Eqs. \eqref{aodd},
1415: \eqref{aeven}), converges to the center and to the edges of the
1416: Brillouin zone and finally collapses to the respective points.
1417: Hence, in the limit of very large $N$ the amplitude of the wave
1418: function decays inside the wire for practically all values of
1419: $k_y$. This means that for $N\to\infty$ the dispersive branch
1420: $\omega_A(k_y)$ transforms into the zero-energy dispersionless
1421: surface states situated near the two surfaces.
1422: 
1423: 
1424: \begin{figure}[tbh]
1425: \begin{center}
1426: \leavevmode
1427: \includegraphics[clip=true,width=.9\columnwidth]{WeightZES.eps}
1428: \caption{The weight of the zero-energy peak in the LDOS, taken at the first layer,
1429: as a function on odd (open circles) and even (filled circles) $N$.}
1430: \label{figweights}
1431: \end{center}
1432: \end{figure}
1433: 
1434: Fig.~\ref{figweights} shows $N$-dependence of the weight of the
1435: zero-energy peak in the LDOS. For odd $N$, the weight diminishes
1436: with increasing $N$, while for even $N$ it increases. In the limit of
1437: large $N$ the two curves will converge to the weight for the
1438: zero-energy surface states when the surfaces are infinitely far
1439: apart. The odd-even effects become negligibly small only for
1440: $Nd\gg \xi_0$. One could try to recover the quasiclassical
1441: results for wide wires $Nd\gg a$ after averaging over odd and even
1442: film widths. The boundary conditions for
1443: quasiclassical propagators are taken somewhere on a distance $l$
1444: from the surface and imply some uncertainty about the boundary
1445: positions, as well as the film (wire) thickness ($a\ll l\ll \xi_0$).
1446: As seen from Fig.~\ref{figweights}, averaging of the weights of the
1447: peaks over odd and even $N$ will strongly reduce their width
1448: dependence. This kind of averaging is much closer (although not
1449: identical) to the quasiclassical results on the width dependence of
1450: the LDOS for the $d$-wave superconducting film \cite{nagai}.  Very
1451: recent quasiclassical results \cite{fominov}, treating various wire
1452: orientations, demonstrate the appearance of energy bands of
1453: quasiparticle states, in particular, for $(110)$ wires. Our
1454: microscopic model for high quality half-filled wires with fixed
1455: number of chains gives, however, only a couple of branches
1456: Eq.\eqref{omegaa} for even $N$ wires and the dispersionless
1457: zero-energy states for odd
1458: $N$ wires. We associate the difference between the microscopic and
1459: the quasiclassical results with the particular condition of
1460: half-filling. The quasiclassical approach implies no singular
1461: behavior of the LDOS in the normal metal state near the Fermi
1462: surface, whereas the Van Hove singularity takes place in the LDOS
1463: on the Fermi surface for the normal metal state of half-filled
1464: infinite square lattice. An agreement of our microscopic results with
1465: the quasiclassical ones arises in the presence of deviations from
1466: half-filling (see below Eqs.\eqref{omegaamu}),\eqref{phi}).
1467: 
1468: As already mentioned, in the even $N$ wires the particle-hole
1469: structure of any quasiparticle states with broken $k_y\to-k_y$
1470: symmetry satisfies the condition $|u(n,k_y)|=|v(n,k_y)|$. Our
1471: picture is that for the states above the gap this Andreev
1472: particle-hole structure is generated by the infinite sequence of
1473: ``overbarrier'' (overgapped) Andreev reflections, induced by a
1474: sign reversal of the order parameter, which the quasiparticles
1475: experience along their trajectories being bounded inside the wire
1476: with impenetrable surfaces. This unconventional feature does not
1477: take place for the states above the gap in the odd $N$ wires,
1478: since the two surfaces always result in the standing waves across
1479: the wire with no important interference effects in this case. For
1480: negligibly small splitting one should consider a superposition of
1481: two wave functions, describing the two split states. Then the
1482: Andreev structure of initially nondegenerate wave functions is
1483: lost, since the moduli of particle and hole amplitudes can easily
1484: differ from each other. The Andreev structure of quasiparticle
1485: wavefunctions can be lost also in the presence of deviations from
1486: the half-filling, if $\mu$ is larger or of the same order as the
1487: splitting. Since the splitting vanishes for $k_y=0$, one can
1488: expect that for sufficiently small $k_y$ the Andreev structure of
1489: the wave functions will be destroyed even for small $\mu$. In the
1490: next section some other consequences of deviations from
1491: half-filling are considered.
1492: 
1493: \subsubsection{Deviations from  half-filling}
1494: 
1495: The shape of the Fermi surface depends on $\mu$ and has a strong
1496: influence on the low-energy quasiparticle spectrum. Consider, for
1497: example, low-energy quasiparticle states under the conditions
1498: $|\omega|,\ |\Delta(k_y)|\ll q(k_y)$. This ensures that the
1499: quasiparticle energies lie  close to the Fermi surface and one can
1500: take their energies in the normal metal state to be in the linear
1501: form $\propto \v_F\cdot(\k-\k_{F})$. Under this approximation,
1502: effects of the particle-hole asymmetry are small and one can use a
1503: quasiclassical approximation, which is valid for quasiparticles
1504: close to the Fermi surface. Then only the order parameter on the
1505: Fermi surface $\Delta(k_{F,x}, k_{F,y})$ enters the equations. For
1506: the wire geometry the momentum component $k_{F,y}$ is an
1507: independent parameter, while $k_{F,x}(k_{F,y})$, and
1508: $\Delta(k_{F,x}(k_{F,y}), k_{F,y})$ are actually functions of
1509: $k_{F,y}$. For $\mu=0$ the Fermi surface for the $(110)$ wire is a
1510: square with sides parallel to $x$ or $y$ axes. Hence, $k_{F,x}=\pm
1511: \pi/(2d)$ actually does not depend on $k_{F,y}$ in this case and
1512: $\Delta(\pm\pi/(2d), k_{F,y})=\pm\Delta(k_{F,y})$, in accordance
1513: with Eq.\eqref{dky}. For finite $\mu$ we find
1514: $\Delta(k_{F,x},k_{F,y})=
1515: \dfrac{\Delta(k_{F,y})}{q(k_{F,y})}\sqrt{q^2(k_{F,y})-\mu^2}$.
1516: 
1517: In the case of superconducting wires the equation for
1518: quasiparticle subgap energies near the Fermi surface
1519: $|\omega|<|\Delta(k_{F,x},k_{F,y})|$ takes the form
1520: \begin{eqnarray} \omega^2\left(
1521: \sinh^2\left[(N+1) d\dfrac{\sqrt{\Delta^2(k_{F,x},
1522: k_{F,y})-\omega^2}}{
1523: |v_{x,f}(k_{F,y})|}\right] \right.\nonumber\\
1524: +\sin^2\phi(k_{F,y})\Biggr)= \Delta^2(k_{F,x},
1525: k_{F,y})\sin^2\phi(k_{F,y})~,\label{spectrumassym}
1526: \end{eqnarray}
1527: where $\phi(k_{F,y}) \equiv k_{F,x}d(N+1)$.
1528:  The lowest branches of quasiparticle spectra, which follow
1529: from Eq. \eqref{spectrumassym} for $N=24,25,26$, are shown in
1530: Fig.\ref{fig:mu02spectrum}. The solution of
1531: Eq.~\eqref{spectrumassym} reduces to a simple form in the limit of
1532: large $N$:
1533: 
1534: \bea \omega_A(k_{F,y})=\pm 2\Delta(k_{F,y})\sin\phi(k_{F,y})\times
1535: \qquad \qquad \qquad
1536: \nonumber\\
1537: \exp\left[-(N+1)d\dfrac{|\Delta(k_{F,y})|}{|v_{x,f}(k_{F,y})|}\right]
1538: \enspace . \label{omegaamu} \eea
1539: 
1540: \begin{figure}[!tbh]
1541: \begin{center}
1542: \leavevmode
1543: \includegraphics[clip=true,width=.9\columnwidth]{StripN24-25-26Mu02.eps}
1544: \caption{The lowest energy branches for N=24, 25, 26. The parameters
1545: are $t=2.5$, $\mu=0.2t=0.5$, $\Delta_0=0.2t=0.5$, $\Delta_{max}=2\Delta_0=1$.}
1546: \label{fig:mu02spectrum}
1547: \end{center}
1548: \end{figure}
1549: 
1550: For half-filled wires, when $\mu=0$, the phase $\phi(k_{F,y})$
1551: does not depend on $k_{F,y}$, being equal to $\phi_{\rm odd}=m\pi$
1552: for odd $N$ wires and $\phi_{\rm even}=(m+1/2)\pi$ for even $N$
1553: wires. This difference between the phases $\phi_{\rm odd}$ and
1554: $\phi_{\rm even}$ plays an important role in forming well
1555: pronounced odd-even effects in the spectra of wires with odd and
1556: even numbers of layers. Indeed, for half-filled odd $N$ wires
1557: Eq.~\eqref{omegaamu} reduces to the zero-energy dispersionless
1558: Andreev states. At the same time, for half-filled even $N$ wires
1559: Eq.~\eqref{omegaamu} describes a dispersive branch of
1560: quasiparticle energies, which coincides with Eq.~\eqref{omegaa}
1561: under the condition $|\Delta(k_y)|\ll q(k_y)$.
1562: 
1563: \begin{figure}[t]
1564: \begin{center}
1565: \leavevmode
1566: \includegraphics[clip=true,width=.9\columnwidth]{mu110.eps}
1567: \caption{Local density of states for a (110) wire
1568: with N=10 (left column) and N=11 (right column); $\mu=0$ (upper panels),
1569: $\mu=0.2t$ (lower panels). On each panel the different curves represent
1570: various chains.}
1571: \label{fig:mu110}
1572: \end{center}
1573: \end{figure}
1574: 
1575: For $\mu\ne0$ the phase $\phi(k_{F,y})$ noticeably depends on
1576: $k_{F,y}$: \be \phi(k_{F,y})=k_{F,x}d(N+1)=
1577: (N+1)\cos^{-1}\left(-\dfrac{\mu}{q(k_{F,y})}\right) \enspace .
1578: \label{phi}\ee
1579: Qualitative deviations of low-energy quasiparticle spectra,
1580: shown in Fig.~\ref{fig:mu02spectrum}, from the respective
1581: spectra of half-filled wires (see
1582: Figs.~\ref{figN11},~\ref{figN10}) are associated with the behavior
1583: of the phase $\phi(k_y)$.
1584: The odd-even effect in the spectra of $(110)$ wires
1585: becomes less pronounced in the case of finite $\mu$, as found by
1586: Ziegler et al. \cite{zieglerwire}. For some values of $\mu$ the
1587: spectra of odd $N$ and even $N$ wires may have no qualitative
1588: differences at all. As it follows from Eq.~\eqref{omegaamu}
1589: and, in a more general case, from Eq.~\eqref{spectrumassym},
1590: for $\mu\ne 0$ additional and  strong dispersion of the spectra
1591: comes from the $k_y$-dependence of the phase $\phi(k_{F,y})$. The
1592: larger the $N$, more oscillations of $\sin\phi(k_y)$ take place
1593: with varying the $k_{F,y}$. Hence, more extrema of
1594: $\omega_A(k_{F,y})$ arise. This results in additional peaks in the
1595: LDOS, which appear only in the presence of finite $\mu$.
1596: As seen from Eq.~\eqref{phi}, the phase $\phi$ can considerably vary,
1597: when the film (wire) thickness  varies from $Nd$ to $Nd-l$ and
1598: $a\ll l\ll\xi_0$, $l\ll Nd$. Thus, in averaging the spectrum over the film
1599: thickness, a large number of respective
1600: additional peaks arises, filling the whole low-energy band
1601: in the $(110)$ wire with edges described by Eq.~\eqref{omegaamu}. This
1602: is in agreement with the quasiclassical results \cite{nagai,fominov}.
1603: 
1604: The particle-hole asymmetry is another important feature of the
1605: spectra. It can be well pronounced for finite $\mu$, but lies
1606: beyond the quasiclassical approximation. Fig.\ref{fig:mu110}
1607: displays the asymmetric LDOS, calculated with Eq.(\ref{11}) for
1608: $(110)$ superconducting wires with $N=10$ and $N=11$ in the case
1609: $\mu=0.2t$. For  comparison, the respective LDOS for $\mu=0$ is
1610: also shown. A well-pronounced peak at $\omega=-\mu$ arises with a
1611: deviation from half-filling in the LDOS for the odd $N$ wires. We
1612: remind the reader that in the half-filled normal state $(110)$
1613: wires with odd $N$ the dispersionless zero-energy quasiparticle
1614: states have been found in Sec.\ref{sec:nwires} (see
1615: Eq.(\ref{modefreqnormal}) with $\nu=(N+1)/2$). The wave function
1616: of these states, as well as the residue of the pole-like term in
1617: the Green's function Eq.(\ref{weightnn}), is a standing wave
1618: across the wire, taking zero values on alternating sites.
1619: In the superconducting state the zero-energy standing wave
1620: disappears and the dispersionless zero-energy Andreev surface
1621: states arise. Their weight exponentially decays in the bulk of
1622: wide odd $N$ wires. In the presence of a deviation from
1623: half-filling the energy of the quasiparticle states in the normal
1624: state odd $N$ wires shifts to $-\mu$. These dispersionless states
1625: with finite energy keep the character of standing waves. Further,
1626: in the superconducting wires with finite $\mu$ the low-energy
1627: states become dispersive both for odd and even $N$
1628: wires. As seen from Fig.(\ref{fig:mu02spectrum})
1629: for the wire with $N=25$, the branch with lowest energy has in
1630: this case two extrema. The maximal value of $|\omega|$ for the
1631: states forming this branch, lies at $k_y=0$ and contributes to the
1632: peak at $\omega=-\mu$ associated with the hole contribution, if
1633: $\mu>0$. Since the order parameter vanishes at $k_y=0$, the
1634: respective quasiparticle wave function is a standing wave and the
1635: peak position coincides with that in the normal metal state of the
1636: wire. The minimal value of $|\omega|$ is zero and contributes to
1637: the zero-energy peak. The zero-energy peak is associated with
1638: comparatively large value of $k_y$, comparable with the size of
1639: the Brillouin zone, and the order parameter at this value of $k_y$
1640: is of order  $\Delta_{max}$. Thus, the zero-energy peak is
1641: associated with surface Andreev states, whose quasiparticle
1642: wave function decays in the bulk of wide odd $N$ wires. For narrow
1643: wires, the self-consistency condition becomes important. As shown
1644: in the next section, at finite $\mu$ the self-consistency
1645: condition can lead to more important consequences as compared with
1646: the case $\mu=0$.
1647: 
1648: 
1649: \label{sec:finitemu}
1650: 
1651: 
1652: 
1653: \section{Self-consistent treatment of $\mathbf{(110)}$-wires}
1654: \label{sec:self}
1655: 
1656: 
1657: In previous sections, the order parameter was assumed constant
1658: over the whole width of the wire in order to allow for analytical
1659: solutions. Even for a single $(110)$-surface, however, a
1660: self-consistent treatment of the order parameter gives rise to
1661: interesting effects: the $d_{x^2-y^2}$--wave order parameter is
1662: strongly suppressed near the surface and a complex $is$--admixture
1663: (or some other time-reversal symmetry-breaking state) is possible, which
1664: leads to a splitting of the zero-energy Andreev bound state~\cite{mats95,%
1665: buch95,cov97,fog97,Ting110,deutsch99,greene03,kos01,sigrist,ohashi99,%
1666: vanHarlingen02,deutsch02,deutsch03}.
1667: Even larger effects are therefore to be expected for the wire limited
1668: by two $(110)$-surfaces.
1669: Indeed, our self-consistent evaluation indicates that for very
1670: narrow wires a quasi one-dimensional triplet superconducting state
1671: can replace the conventional $d_{x^2-y^2}+is$--state. For finite
1672: chemical potential
1673: $\mu$ the normal metal state can become energetically favorable
1674: as the ground state for narrow wires, while superconductivity recovers
1675: with increasing wire width. Under special
1676: conditions, even a mixture of singlet- and triplet-pairing can
1677: occur.
1678: It is important to recall at this point that one
1679: expects mean field theory to break down as the 1D limit is
1680: approached even at $T=0$.   Thus the predictions for various kinds
1681: of superconducting order mentioned below are to be treated with
1682: some skepticism as regards quantitative predictions.  Nevertheless
1683: we view our results as presenting intriguing evidence that when
1684: surface energies begin to become comparable to the energy
1685: differences between $d$--wave and other bulk pair channels, strong
1686: fluctuations with symmetries optimal for quasi-1D system,
1687: including spin triplet pair fluctuations, will result.
1688: 
1689: 
1690: 
1691: \subsection{Order parameter}
1692: 
1693: Self-consistent solutions to Hamiltonian~(\ref{ham}) were
1694: obtained by solving the Bogoliubov-de Gennes equations for
1695: $(110)$-wires,
1696: \begin{eqnarray}
1697: \left ( \begin{array}{cc} \xi_{k_y} & \Delta_{k_y} \\
1698:                           \Delta_{-k_y}^* & -\xi_{k_y}
1699:         \end{array} \right)
1700: \left ( u_{k_y\lambda} \atop{v_{k_y\lambda}} \right ) =
1701: E_{k_y\lambda} \left ( u_{k_y\lambda} \atop{v_{k_y\lambda}} \right)
1702: \end{eqnarray}
1703: where $E_{k_y\lambda}$ with $\lambda = 1 \ldots N$
1704: are the eigenvalues of the Bogoliubov-de Gennes
1705: equations and its eigenvectors $u_{k_y\lambda}(n)$ and
1706: $v_{k_y\lambda}(n)$ are the
1707: coefficients of the Bogoliubov transformation:
1708: \begin{eqnarray}
1709: c_{k_yn\uparrow}&=&\sum_\lambda \left \{
1710:                  \gamma_{k_y \lambda \uparrow} u_{k_y \lambda} (n)
1711:            - \gamma_{k_y \lambda \downarrow}^\dag v^*_{k_y \lambda} (n)
1712:         \right \}
1713: \nonumber \\
1714: c_{-k_yn\downarrow}^\dag&=&\sum_\lambda \left \{
1715:                  \gamma_{k_y \lambda \uparrow} v_{k_y \lambda} (n)
1716:            + \gamma_{k_y \lambda \downarrow}^\dag u^*_{k_y \lambda} (n)
1717:         \right \}
1718: \end{eqnarray}
1719: Furthermore, $\xi_{k_y}$ and $\Delta_{k_y}$ are matrices in the $N$
1720: layers of the $(110)$-wire, i.e. the layers in $x$-direction, and are
1721: given by (for notation see Fig.~\ref{fig:strip110}):
1722: \begin{eqnarray}
1723: (\xi_{k_y})_{nn'} &=& -2t \cos k_y \left(
1724: \delta_{n'n+1}+\delta_{n'n-1}
1725:                                           \right)  -\mu \delta_{n'n}
1726: \nonumber \\
1727: \left (\Delta_{k_y} \right)_{nn'} &=&
1728:       \left( \Delta_{nn'}^+e^{-ik_y}+\Delta_{nn'}^-e^{ik_y} \right)
1729:                                                          \delta_{n'n+1}
1730: \nonumber \\
1731: &+&   \left( \Delta_{nn'}^-e^{-ik_y}+\Delta_{nn'}^+e^{ik_y}
1732: \right)
1733:                                                          \delta_{n'n-1}
1734: \end{eqnarray}
1735: The gap values are determined by the following self-consistency
1736: equations:
1737: \begin{eqnarray}
1738: \Delta_{nn+1}^\pm &=& - V \frac{1}{N} \sum_{k_y} e^{\pm ik_y}
1739:           \langle c_{-k_y n+1 \downarrow} c_{k_y n \uparrow} \rangle
1740: \\
1741:                &=& V \frac{1}{N} \sum_{k_y} \sum_\lambda e^{\pm ik_y}
1742:           u_{k_y\lambda}(n+1) v_{k_y\lambda}^*(n)
1743: \nonumber \label{eq:GapSC}
1744: \end{eqnarray}
1745: 
1746: 
1747: 
1748: To simplify the numerical evaluation of the Bogoliubov-de Gennes
1749: equations we consider isolated wires here. For a Hamiltonian on a
1750: discrete lattice like~(\ref{ham}) an isolated wire is equivalent
1751: to a wire limited by lines of unitary impurities, which is the
1752: boundary condition used for the analytical calculations in the
1753: previous sections.
1754: 
1755: \begin{figure}[t]
1756: \begin{center}
1757: \includegraphics*[width=.99\columnwidth]{DdsmaxD01.eps}
1758: \\[.5cm]
1759: \includegraphics*[width=.99\columnwidth]{PhaseDiagD01.eps}
1760: \end{center}
1761: \caption{Narrow wires with a nearest neighbor interaction strength of
1762: $V$=1.1575$t$, which gives rise to a gap value of $\Delta_0$=0.2$t$ for
1763: the bulk system at $\mu$=0.
1764: Lower panels: phase diagram of the wires with even width
1765: (c) and with odd width (d). White space
1766: denotes $\Delta=0$, i.e., the normal state.
1767: Upper panels: corresponding amplitudes of the $d_{x^2-y^2}$ and $is$-order
1768: parameters displaying the value in the center of the wire for
1769: the $d_{x^2-y^2}$--wave case and the value at the edge of the
1770: wire in the $s$--wave case. These are the positions where the largest
1771: values of the respective order parameters are expected in the
1772: usual $d_{x^2-y^2}+is$-state.}
1773: \label{fig:PhaseDiagD01}
1774: \end{figure}
1775: 
1776: 
1777: 
1778: For narrow wires we find a variety of different phases.
1779: The resulting phase diagram for a nearest neighbor
1780: interaction strength of $V$=1.1575$t$, which gives rise to a gap
1781: of $\Delta_0$=0.2$t$ in a bulk system at $\mu$=0, is displayed in
1782: Fig.~\ref{fig:PhaseDiagD01}. For wires with widths up to $N$=9 we
1783: find a new phase over a wide range of chemical potentials
1784: characterized by $\Delta_{ij}=-\Delta_{ji}$ (stars in
1785: Fig.~\ref{fig:PhaseDiagD01}(c), (d)), which is a signature of
1786: triplet-pairing with $S_z=0$, the only triplet component
1787: compatible with Hamiltonian~(\ref{ham}). For $\mu=0$ the amplitude of
1788: $\Delta_{ij}$ oscillates across the wire between zero and its
1789: maximum value (see Fig.~\ref{fig:DensN08}(a)), indicating a
1790: one-dimensional nature of these new triplet-superconducting
1791: correlations (see Fig.~\ref{fig:DensN08}(b)).
1792: Although the oscillating behavior of $\Delta_{ij}$
1793: remains for finite $\mu$, its amplitude no longer vanishes exactly on
1794: alternating layers. Thus the strict one-dimensionality of the superconducting
1795: correlations seems to be a feature peculiar to $\mu=0$.
1796: This new triplet superconducting phase
1797: will be discussed in more detail below.
1798: At first we will focus on the singlet superconducting phase with possible
1799: $d_{x^2-y^2}$-- and $s$--wave order parameters.
1800: 
1801: 
1802: \begin{figure}[t]
1803: \begin{center}
1804: \begin{minipage}{.49\columnwidth}
1805: \includegraphics*[width=.98\columnwidth,clip=true]{TripletStateN6.eps}
1806: \\[.3cm]
1807: \includegraphics*[width=.98\columnwidth,clip=true]{DensN08p2.eps}
1808: \end{minipage}
1809: \begin{minipage}{.49\columnwidth}
1810: \includegraphics*[width=.98\columnwidth,clip=true]{DensN08p1.eps}
1811: \end{minipage}
1812: \end{center}
1813: \caption{Quasi one-dimensional triplet superconducting state for $N$=8 and
1814: $\mu$=0. (a) Bond gap values $\Delta^\pm_{nn+1}$ (for definition see
1815: Eq.~(\ref{eq:GapSC})) across the wire.
1816: (b) Illustration of the one-dimensional structure of the
1817: superconducting correlations for an $N$=6-(110) wire at $\mu=0$.
1818: (c) Density of states starting from the outermost
1819: layer of the wire up to the middle layer of the wire, where for
1820: illustrational purposes each
1821: layer has been shifted by an additional offset of 0.3$t$.}
1822: \label{fig:DensN08}
1823: \end{figure}
1824: 
1825: 
1826: Although we do not understand all details of the variability of
1827: the phase diagram, a few general trends seem clear.
1828: The larger the
1829: width of the wire the more of the usual
1830: $d_{x^2-y^2}+is$--phase (filled circles in
1831: Fig.~\ref{fig:PhaseDiagD01}(c),(d)) is
1832: recovered. The corresponding amplitudes of
1833: the $d_{x^2-y^2}$-- and $is$--components of the order
1834: parameter are displayed in the upper panels of
1835: Fig.~\ref{fig:PhaseDiagD01}. For narrow wires the amplitude of the
1836: $d_{x^2-y^2}$--wave order parameter is finite only for small chemical
1837: potentials $\mu$ and a pure $s$--wave phase
1838: is favorable for large $\mu$. These two phases are separated by a
1839: normal state region.
1840: For larger wire widths the range of the $d_{x^2-y^2}$--wave phase
1841: and the amplitude of the $d_{x^2-y^2}$--wave order parameter
1842: increase. The amplitude of the $s$--wave order parameter, on the other
1843: hand, decreases and the $s$--wave phase moves
1844: towards smaller chemical potentials until it merges with the
1845: $d_{x^2-y^2}$--wave phase, thereby giving rise to a finite
1846: $is$--admixture near the edges of the wire, as expected in
1847: analogy to a single (110)-surface.
1848: 
1849: 
1850: The dependence of the amplitude of the $d_{x^2-y^2}$--wave order
1851: parameter on the width $N$ of the wire is shown in
1852: Fig.~\ref{fig:D01D02}. The upper panels refer to an interaction
1853: strength  of $V$=1.1575$t$, which gives
1854: rise to a gap of $\Delta_0$=0.2$t$ for the bulk system at $\mu$=0.
1855: For $\mu$=0 we observe an even-odd
1856: oscillation in the amplitude of the order parameter (see
1857: Fig.~\ref{fig:D01D02}(a)), which disappears for
1858: larger chemical potentials. The effect of finite $\mu$ on the suppression
1859: of the $d_{x^2-y^2}$--wave order parameter is
1860: considerably stronger in narrow wires than in the bulk system.
1861: For $\mu/t$=0.4 the $d_{x^2-y^2}$--wave order parameter vanishes for
1862: $N<10$ (squares in Fig.~\ref{fig:D01D02}(a)), although the bulk gap
1863: is only reduced by a factor of approx. 0.8 (see
1864: Fig.~\ref{fig:D01D02}(c)).
1865: The dramatic suppression of the $d_{x^2-y^2}$--wave order parameter
1866: for finite $\mu$ is reduced upon consideration of a larger nearest neighbor interaction
1867: strengths, as can be seen in the lower panels of Fig.~\ref{fig:D01D02}.
1868: For a larger interaction strength of $V$=1.7682$t$, which gives rise to
1869: a bulk gap value of
1870: $\Delta_0$=0.4$t$ at $\mu=0$, a pronounced even-odd effect remains at
1871: $\mu/t$=0.4 (see Fig.~\ref{fig:D01D02}(d)) and the
1872: amplitude in the center of an $N$=12-wire is already very close to the bulk
1873: value (see Fig.~\ref{fig:D01D02}(f)) contrary to the smaller
1874: interaction strength, where the amplitude in the center of an
1875: $N$=12-wire is still suppressed by a factor of more than three with
1876: respect to the bulk value (see Fig.~\ref{fig:D01D02}(c)).
1877: 
1878: 
1879: \begin{figure}[t]
1880: \begin{center}
1881: \includegraphics*[width=.99\columnwidth]{D01D02.eps}
1882: \end{center}
1883: \caption{Evolution of the $d_{x^2-y^2}$-wave order parameter with
1884: increasing wire width for two different interaction strengths
1885: $V$=1.1575$t$ (upper panels) and $V$=1.7682$t$ (lower panels), which
1886: give rise to a gap value of $\Delta_0$=0.2$t$ and $\Delta_0$=0.4$t$  for the
1887: bulk system at $\mu=0$, respectively. The first column displays
1888: the amplitude of the $d_{x^2-y^2}$-order parameter as a function of the
1889: wire width for two different chemical potentials $\mu/t$=0 and $\mu/t=0.4$.
1890: The other two columns show the variation of the $d_{x^2-y^2}$- and
1891: $is$-component of the order parameter across the wire for $\mu/t$=0
1892: (center column) and $\mu/t$=0.4 (right column). For comparison the
1893: reduced magnitude of the bulk order parameter at $\mu/t$=0.4
1894: is displayed with dashed lines in the right panels.}
1895: \label{fig:D01D02}
1896: \end{figure}
1897: 
1898: 
1899: 
1900: 
1901: The emergence of a new triplet-superconducting phase for very
1902: narrow wires can be most easily understood by considering  the
1903: smallest wire, i.e., the $N$=2 wire. Although the effect of
1904: fluctuations will be larger for smaller wires, we focus only on
1905: possible solutions of the BCS mean-field Hamiltonian~(\ref{ham})
1906: in the present paper. Inspection of the phase diagram of
1907: Fig.~\ref{fig:PhaseDiagD01} shows that the $d$-wave order
1908: parameter vanishes for the $N$=2-wire. To investigate alternative
1909: ways in which the $N$=2 wire could lower its ground state energy,
1910: we map it to a 1D-chain (see Fig.~\ref{fig:wire1d}). Although the
1911: gap values $\Delta_{ij}$ and $\Delta_{ji}$ could in principle
1912: differ in amplitude and by a phase factor $\phi$, we restrict our
1913: treatment to $\phi=0$ and $\phi=\pi$. Note that $\phi=0$
1914: corresponds to a singlet pairing state, whereas $\phi=\pi$ is a
1915: triplet pairing state with $S_z=0$\cite{tripletfootnote}. To
1916: investigate the possibility of triplet superconductivity in these
1917: systems in more generality, one should retain pairing correlations
1918: with $S_z=\pm 1$ in the mean-field Hamiltonian as well. For now,
1919: however, we are satisfied with the observation that even for the
1920: simple nearest-neighbor pairing Hamiltonian~(\ref{ham}) a triplet
1921: order parameter can be favored over a singlet order parameter in
1922: narrow geometries.
1923: 
1924: \begin{figure}[h]
1925: \begin{center}
1926: \includegraphics*[width=.99\columnwidth]{strip1d.eps}
1927: \end{center}
1928: \caption{Mapping of the $N$=2 wire with $(110)$ orientation to
1929: the 1D-chain with lattice constant $d$. Note that
1930: $\Delta_2=-\Delta_1$ corresponds to a triplet pairing state with
1931: $S_z$=0.} \label{fig:wire1d}
1932: \end{figure}
1933: 
1934: 
1935: After Fourier transformation Hamiltonian~(\ref{ham}) for the
1936: 1D-chain reads
1937: \begin{eqnarray}
1938: H_{\rm MF}&=&-\sum_{k \sigma} (2t \cos kd + \mu)
1939:                             c_{k \sigma}^\dag c_{k \sigma}
1940: \\
1941:        & &    +\sum_k \left \{ (\Delta_1 e^{ikd}+\Delta_2 e^{-ikd})
1942:                             c^\dagger_{-k \downarrow} c^\dagger_{k \uparrow}
1943:                             + {\rm h.c.} \right\}\,, \nonumber
1944: \label{eq:Ham1d}
1945: \end{eqnarray}
1946: where $\Delta_{ii+1}=-V\langle c_{i+1\downarrow}
1947: c_{i\uparrow}\rangle=\Delta_1$ and $\Delta_{i+1i}=-V\langle
1948: c_{i\downarrow} c_{i+1\uparrow}\rangle=\Delta_2$. It can be easily
1949: diagonalized using the
1950: Bogoliubov-transformation
1951: to give a quasiparticle dispersion of:
1952: \begin{eqnarray}
1953: &E_k&=\sqrt{\epsilon_k^2+ \Delta_k^2} \\
1954: &{\rm with}&
1955: \epsilon_k = \frac{q(k)}{2}+\mu=2t \cos kd + \mu
1956: \nonumber\\
1957: & &\!\!\Delta_k^2 =
1958: \Delta_1^2+\Delta_2^2 +2 \Delta_1 \Delta_2 \cos2kd \,, \nonumber
1959: \end{eqnarray}
1960: where the gap values are to be determined from the following
1961: self-consistency  equations:
1962: \begin{eqnarray}
1963: \Delta_1 &=& - V \frac{1}{4\pi d} \int_{-\pi d}^{\pi d} dk
1964:                     \frac{\Delta_1 + \Delta_2 \cos 2kd}{E_k}
1965: \nonumber \\
1966: \Delta_2 &=& - V \frac{1}{4\pi d} \int_{-\pi d}^{\pi d} dk
1967:                     \frac{\Delta_1 \cos 2kd + \Delta_2}{E_k} \,.
1968: \end{eqnarray}
1969: Four different phases emerge from this model by variation of the
1970: chemical potential $\mu$ and the nearest neighbor interaction
1971: strength $V$ (see Fig.~\ref{fig:PhaseDiag1d}). Below a critical
1972: interaction strength the order parameter vanishes and
1973: the normal state is the ground state of the 1D chain. For
1974: intermediate interaction strengths and small chemical potentials
1975: we find $\Delta_1=-\Delta_2$, i.e., a pure triplet superconducting
1976: state, whereas for large chemical potentials the ground state is
1977: characterized by $\Delta_1=\Delta_2$, i.e., a singlet extended
1978: $s$-wave state. In the singlet state the gap function is $\Delta_k
1979: \sim \cos kd$, i.e., it has nodes at $kd=\pm \pi/2$ whereas the
1980: triplet state gap function is $\Delta_k \sim \sin kd$ and therefore
1981: is maximum at $kd=\pm \pi/2$. This explains why the triplet state
1982: is favored over the singlet state for $\mu=0$, where the Fermi-surface
1983: of the nearest neighbor tight-binding model is at $kd=\pm \pi/2$.
1984: For larger chemical potentials the situation is reversed and the
1985: singlet extended $s$-wave state becomes more favorable than the
1986: triplet state.
1987: For large interaction strengths, we find different magnitudes for
1988: $\Delta_1$ and $\Delta_2$, which corresponds to a mixture of
1989: singlet and triplet pairing. In the limit $V \to \infty$, either
1990: of the gap values $\Delta_1$ and $\Delta_2$ approaches zero
1991: whereas the other goes to infinity, corresponding to an admixture
1992: of triplet and singlet order parameters with equal amplitudes.
1993: 
1994: \begin{figure}[t!]
1995: \begin{center}
1996: \includegraphics*[width=.85\columnwidth,clip=true]{PhaseDiag1d.eps}
1997: \end{center}
1998: \caption{Phase diagram for 1D-chain as a function of chemical
1999: potential $\mu/t$ and nearest neighbor interaction $V/t$.}
2000: \label{fig:PhaseDiag1d}
2001: \end{figure}
2002: 
2003: 
2004: With the phase diagram of the 1D-chain in mind we are now able to
2005: better understand the phase diagram of narrow wires as displayed
2006: in Fig.~\ref{fig:PhaseDiagD01}. For small $\mu$ the new phase with
2007: $\Delta_{ij}=-\Delta_{ji}$ simply arises from the formation of
2008: quasi one-dimensional triplet-superconducting correlations along $N$=2
2009: wires (see Fig.~\ref{fig:DensN08}(a),(b)).
2010: This also explains why this new phase is more favorable
2011: for wires with even width than with odd width (compare left and right panels in
2012: Fig.~\ref{fig:PhaseDiagD01}), as only the former can be
2013: divided evenly into $N$=2 wires.
2014: Note that also for the narrow wires,
2015: a critical coupling strength of similar magnitude as in the
2016: 1D-chain is necessary to induce the triplet superconducting state,
2017: which is, however, smaller than the
2018: interaction strength considered in this paper.
2019: At larger chemical potentials we find an extended $s$-wave order
2020: parameter analogously to the one-dimensional chain. More difficult
2021: to reconcile with the phase diagram of the one-dimensional chain,
2022: however, are  the seemingly arbitrarily
2023: distributed mixtures of singlet- and triplet-superconducting order
2024: parameters (crosses in Fig.~\ref{fig:PhaseDiagD01}).
2025: The normal state regions, which occur for narrow wires and
2026: intermediate chemical potentials in Fig.~\ref{fig:PhaseDiagD01},
2027: reflect the fact that superconductivity becomes less favorable in
2028: finite geometries and a critical coupling strength is necessary to
2029: induce it \cite{nagai}.
2030: 
2031: 
2032: 
2033: \subsection{Density of states}
2034: 
2035: 
2036: 
2037: 
2038: 
2039: 
2040: \begin{figure}[t]
2041: \begin{center}
2042: \includegraphics*[width=.99\columnwidth,clip=true]{DensN10N11.eps}
2043: \end{center}
2044: \caption{Density of states for the $N$=10-wire (left panels) and
2045: the $N$=11-wire (right panels). Upper panels show the
2046: non-selfconsistent results assuming a constant $d_{x^2-y^2}$-order
2047: parameter of $\Delta_0$=0.2$t$, whereas the lower panels display
2048: the density of states resulting from the self-consistently
2049: determined values of the $d$-wave order parameter, whose variation
2050: throughout the wire is depicted in the insets. All panels show the
2051: density of states starting from the outermost layer of the wire up
2052: to the middle layer of the wire, where for each layer towards the
2053: center of the wire an additional offset of 0.3$t$ has been used.}
2054: \label{fig:DensN10N11}
2055: \end{figure}
2056: 
2057: 
2058: 
2059: 
2060: What does the phase diagram in Fig.~\ref{fig:PhaseDiagD01} imply
2061: about the possible existence of Andreev bound states in narrow
2062: wires? In Fig.~\ref{fig:DensN08}(c) the local density of states and
2063: the variation of the order parameter throughout the wire is
2064: depicted for the quasi one-dimensional triplet superconducting
2065: state using the $N$=8-wire at $\mu$=0 as an example. Obviously,
2066: the density of states in this state is fully gapped, in analogy to
2067: the one-dimensional chain, and there are no Andreev bound states.
2068: 
2069: Close to half-filling, for wider wire widths, the results are more
2070: conventional and the effects of self-consistency much simpler.  In
2071: Fig.~\ref{fig:DensN10N11} the local density of states is displayed
2072: for the $N$=10- and the $N$=11-wire at $\mu=0$, which are
2073: characterized by a pure $d_{x^2-y^2}$-order parameter. Here, the
2074: main difference between the self-consistent (lower panels of
2075: Fig.~\ref{fig:DensN10N11}) and the non-self consistent results
2076: (upper panels of Fig.~\ref{fig:DensN10N11}) is the suppression of
2077: the magnitude of the $d$-wave order parameter especially towards
2078: the edges of the wire (see also insets of
2079: Fig.~\ref{fig:DensN10N11}). Whereas the  weight of the zero-energy
2080: peak is reduced in the even-width ($N$=10) wire the weight of
2081: the zero-energy state in the odd-width wire ($N$=11)
2082: is hardly affected.
2083: 
2084: 
2085: 
2086: 
2087: 
2088: 
2089: 
2090: 
2091: 
2092: 
2093: 
2094: 
2095: 
2096: \section{conclusion}
2097: 
2098: \label{sec:conclusion}
2099: 
2100: Motivated by recent scanning tunneling experiments on the
2101: Ba$_2$Sr$_2$CaCu$_2$O$_8$ systems which reveal inhomogeneous
2102: electronic structure on the nanoscale, we have analyzed in detail
2103: the electronic structure of $d$-wave quantum wires. These wires
2104: exemplify the effects of a constrained geometry while still being
2105: simple enough to allow for analytical solutions. To impose a
2106: restricted geometry we use lines of impurities with infinite
2107: scattering strength, a method which allows to cut arbitrarily
2108: shaped objects out of the two-dimensional plane. In principle, it
2109: is straightforward to extend this method to investigate the
2110: effects of tunneling between neighboring grains by reducing the
2111: scattering strength of the impurities and thus lowering the
2112: potential barrier between neighboring grains.
2113: 
2114: New and interesting physics arises from interference effects
2115: between the two surfaces of the wire when its width is of the
2116: order of the superconducting coherence length. Contrary to
2117: $s$-wave superconductors in finite geometries, the surface pair
2118: breaking plays an important role. In this respect, the existence
2119: and nature of Andreev bound states in constricted geometry is of
2120: particular interest. In order to single out new effects peculiar
2121: to quantum wires and arising from the interference of the two
2122: surfaces, we have first addressed the case of a single surface in
2123: the first part of the paper, concentrating on surfaces with
2124: $(100)$-, $(210)$- and $(110)$-orientations in a nearest neighbor
2125: tight-binding model at half-filling. Andreev bound states can form
2126: on surfaces with orientations deviating from the $(100)$ direction
2127: due to the sign change of the $d$-wave order parameter. However, in
2128: the presence of several channels for reflection of quasiparticles
2129: from the surface, the zero-energy Andreev states may not exist, as
2130:  is the case for the $(210)$ surface of the square lattice. Here
2131: our results are in qualitative agreement with earlier work based
2132: either on the quasiclassical approximation or Bogoliubov-de Gennes
2133: equations. Only for the $(110)$-surface, which involves the
2134: strongest pair-breaking, do we find a zero-energy Andreev bound
2135: state, whose amplitude decreases with the square of the inverse
2136: distance from the surface and vanishes on even layers.
2137: 
2138: In the main part of this work, we  focussed on quantum wires with
2139: $(110)$-orientation, which
2140: display a pronounced ``width parity" effect.  The special case of
2141: electrons hopping on a square lattice with half-filled
2142: tight-binding band was treated most extensively.  For wires of
2143: this type with an odd number of chains $N$ only one subgap state,
2144: a dispersionless zero-energy Andreev-bound state, exists. As in
2145: the single surface case, the amplitude of the ZES vanishes on even
2146: layers and decays towards the center of the wire. In addition to
2147: the zero-energy state (ZES), there are $N$-1 dispersive modes,
2148: which are doubly degenerate due to particle-hole symmetry and are
2149: not Andreev states.
2150: 
2151: In the case of even $N$ half-filled wires, a splitting of the
2152: branches occurs which is associated with a symmetry breaking $k_y
2153: \to - k_y$, and a total of 2$N$ dispersive modes exist. Although
2154: there is no dispersionless ZES,  all quasiparticle states in the
2155: even width wire are of Andreev character in the sense that the
2156: current of the probability density vanishes due to opposite
2157: contributions from particle and hole excitations.  These
2158: quasiparticles occupy either conventional Andreev-type surface
2159: states, or a new type of Andreev standing wave, according to their
2160: momentum $k_y$ parallel to the wire. In the 2D limit $N \to
2161: \infty$,  the lowest energy dispersive state was shown to
2162: transform into the usual zero-energy Andreev bound state at the
2163: impenetrable surface, while the standing wave states evolve
2164: into either the continuous spectrum or the surface states. With
2165: increasing deviation from  half-filling, odd-even effects in
2166: the wires become less pronounced. The evolution of the Fermi surface
2167: shape with these deviations can result in additional extrema in
2168: the quasiparticle dispersive modes and, hence, new peaks in the
2169: LDOS. Large-scale faceting of the surfaces with characteristic
2170: scales larger than the wire thickness will not influence our
2171: results significantly. However, small-scale inhomogeneities like point-like
2172: defects and impurities can substantially change the effects of
2173: interference induced by wire surfaces, even if the phase breaking
2174: length is large.
2175: 
2176: The small coherence length of high-temperature superconductors
2177: leads to further restrictions on the applicability of the
2178: quasiclassical results to superconducting wires or films, whose width is
2179: less than or comparable to the coherence length. For narrow wires of
2180: $(110)$-orientation, the self-consistent treatment of the
2181: order-parameter is found to have a large effect. For wires with
2182: widths less than the superconducting coherence length (up to $N$=9
2183: for the particular parameters used in our numerical calculations),
2184: especially for the even width wires, a new phase characterized by
2185: quasi one-dimensional triplet pairing is found in the mean-field
2186: phase diagram. This new phase is fully gapped and characterized by
2187: the absence of Andreev bound states. The larger the width of the
2188: strip, the more the $d+is$-state, which is expected for a single
2189: surface with $(110)$-orientation, dominates the phase diagram near
2190: half filling. With respect to the density of states at
2191: half-filling, where the order parameter has only a $d$-wave
2192: component, the main effect of the self-consistent treatment is the
2193: suppression of the magnitude of the $d$-wave order parameter,
2194: especially near the surfaces of the wire.
2195: 
2196: It is interesting to end this discussion with some speculations on
2197: the role of bound quasiparticle states and edge effects of this
2198: type on the spectra of weakly coupled superconducting grains as
2199: apparently observed in STM experiments. Such irregular grains
2200: should contain nanoscale ``facets" at all possible angles, so
2201: presumably the most general situation with sizeable particle-hole
2202: asymmetry and mixture of even- and odd-$N$ boundary conditions
2203: will apply. If we first assume that the pair interaction and grain
2204: size are such that one may ignore the triplet states found in the
2205: self-consistent treatment, we expect the spectra of weakly coupled
2206: grains to be dominated by the zero-dimensional analogs of the
2207: dispersive subgap states discussed here, i.e. there should be a
2208: wide distribution of bound state energies depending on local
2209: geometry of the grain, and visible in the LDOS as measured by STM.
2210: In this sense we question whether the weakly-coupled grain picture
2211: is in fact applicable to the experiments in question, which appear
2212: to see a very {\it homogeneous} spectrum  at low energies in the
2213: superconducting state.
2214: 
2215: A second remark is based on our observation that in nanoscale
2216: confined geometry, spin triplet fluctuations may become more
2217: favorable. Such time-reversal symmetry breaking fluctuations will
2218: clearly lead to local spontaneous currents, an issue which has
2219: recently been raised again in angle-resolved photoemission
2220: studies\cite{CampuzanoTbreaking}. Future studies of small grains
2221: are planned to address these issues.
2222: 
2223: \section{Acknowledgments}
2224: We thank I.~Bobkova, Ya.~Fominov, T.~Kopp, J.~Mannhart and K.~Ziegler for useful
2225: discussions. This work was supported, in part, by grant RFBR 02-02-16643
2226: (A.M.B. and Yu.S.B.) and NSF-INT-0340536 (P.J.H. and Yu.S.B). A.M.B.
2227: acknowledges the support of Dynasty Foundation and T.S.N.
2228: the support of the Alexander von Humboldt Foundation.
2229: 
2230: 
2231: \begin{thebibliography}{stuffstuffstuff}
2232: 
2233: \bibitem{Hess91} H.F. Hess, R.B. Robinson, R.C. Dynes, J.M. Valles,
2234: and J.V. Waszak, Phys. Rev. Lett. {\bf 62}, 214 (1989).
2235: 
2236: \bibitem{Fischer95} I. Maggio-Aprile, Ch. Renner, A. Erb, E. Walker,
2237: and \O. Fischer, Phys. Rev. Lett. {\bf 75}, 2754 (1995).
2238: 
2239: \bibitem{yazdani} Ali Yazdani, C. M. Howald, C. P. Lutz,
2240: A. Kapitulnik, and D. M. Eigler, Phys. Rev. Lett.  {\bf 83}, 176
2241: (1999).
2242: 
2243: \bibitem{davisnative} E.W. Hudson, S.H. Pan, A.K. Gupta, K-W Ng, and J.C.
2244: Davis, Science {\bf 285}, 88 (1999)
2245: 
2246: \bibitem{davisZn}
2247: S. H. Pan, E. W. Hudson, K. M. Lang, H. Eisaki, S. Uchida, J. C. Davis,
2248: Nature, 403, 746 (2000).
2249: 
2250: 
2251: \bibitem{cren} T. Cren {\em et al.}, Phys. Rev. Lett. {\bf 84}, 147 (2000).
2252: 
2253: \bibitem{davisinhom1} S.-H. Pan et al., Nature
2254: {\bf 413}, 282 (2001).
2255: 
2256: \bibitem{delozanne} D. J. Derro, E. W. Hudson, K. M. Lang, S. H. Pan,
2257: J. C. Davis, J. T. Markert, and A. L. de Lozanne, Phys. Rev. Lett.
2258: 88, 097002 (2002).
2259: 
2260: \bibitem{howald} C. Howald et al., Phys. Rev. B 64, 100504 (2001).
2261: 
2262: \bibitem{davisinhom2} K.M. Lang et al., Nature 415, 412 (2002).
2263: 
2264: \bibitem{vonDelftreview}
2265: J. von Delft, D. C. Ralph, Phys. Rep. 345, 61 (2001).
2266: 
2267: \bibitem{Tinkham}
2268: C. N. Lau, N. Markovic, M. Bockrath, A. Bezryadin, and M. Tinkham,
2269: Phys. Rev. Lett. 87, 217003 (2001).
2270: 
2271: \bibitem{zieglerwire} K. Ziegler, W.A. Atkinson and P.J.
2272: Hirschfeld, Phys. Rev. B 64, 054512 (2001).
2273: %
2274: \bibitem{geerk88}
2275: J.~Geerk, X.~Xi and G.~Linker,
2276: Z.~Phys. B {\bf 73}, 329 (1988).
2277: %
2278: \bibitem{hu94}
2279: C.-R.~Hu, Phys. Rev. Lett. {\bf 72}, 1526 (1994).
2280: %
2281: \bibitem{tan95}
2282: Y.~Tanaka and S.~Kashiwaya,
2283: Phys.~Rev.~Lett. {\bf 74}, 3451 (1995).
2284: %
2285: \bibitem{mats95}
2286: M.~Matsumoto and H.~Shiba,
2287: J.~Phys.~Soc.~Jpn. {\bf 64}, 1703 (1995);
2288: {\bf 64}, 3384 (1995); {\bf 64}, 4867 (1995); {\bf 65}, 2194 (1996).
2289: %
2290: \bibitem{buch95}
2291: L.~J.~Buchholtz, M.~Palumbo, D.~Rainer, and J.~A.~Sauls,
2292: J. Low Temp. Phys. {\bf 101}, 1079 (1995); {\bf 101}, 1099 (1995).
2293: %
2294: \bibitem{tan952}
2295: S.~Kashiwaya, Y.~Tanaka, M.~Koyanagi, H.~Takashima and K.~Kajumura,
2296: Phys. Rev. B {\bf 51}, 1350 (1995).
2297: %
2298: \bibitem{cov96}
2299: M.~Covington, R.~Scheuerer, K.~Bloom, and L.~H.~Greene,
2300: Appl.~Phys.~Lett. {\bf 68}, 1717 (1996).
2301: %
2302: \bibitem{xu96}
2303: J.~H.~Xu, J.~H.~Miller, Jr., and C.~S.~Ting,
2304: Phys. Rev. B {\bf 53}, 3604 (1996).
2305: %
2306: \bibitem{fog97}
2307: M.~Fogelstr\"om, D.~Rainer, and J.~A.~Sauls,
2308: Phys.~Rev.~Lett. {\bf 79}, 281 (1997).
2309: %
2310: \bibitem{bbs97} Yu.~Barash, H.~Burkhardt, A.~Svidzinsky,
2311: Phys. Rev. B {\bf 55}, 15282 (1997).
2312: %
2313: \bibitem{cov97}
2314: M.~Covington, M.~Aprili, E.~Paraoanu, L.~H.~Greene, F.~Xu, J.~Zhu, and
2315: C.~A.~Mirkin, Phys.~Rev.~Lett. {\bf 79}, 277 (1997).
2316: %
2317: \bibitem{alf97}
2318: L.~Alff, H.~Takashima, S.~Kashiwaya, N.~Terada, H.~Ihara, Y.~Tanaka,
2319: M.~Koyanagi, and K.~Kajimura,  Phys. Rev. B {\bf 55}, R14757 (1997).
2320: %
2321: \bibitem{ek97}
2322: J.~W.~Ekin, Y.~Xu, S.~Mao, T.~Venkatesan, D.~W.~Face, M.~Eddy,
2323: and S.~A.~Wolf,
2324: Phys.~Rev.~B {\bf 56}, 13746 (1997).
2325: %
2326: \bibitem{ap98}
2327: M.~Aprili, M.~Covington, E.~Paraoanu, B.~Niedermeier, and L.~H.~Greene,
2328: Phys.~Rev.~B {\bf 57}, R8139 (1998).
2329: %
2330: \bibitem{alf981}
2331: L.~Alff, S.~Kleefisch, U.Schoop, M.~Zittartz, T.~Kemen, T.~Bauch,
2332: A.~Marx , and R.~Gross,
2333: Eur.~Phys.~J. B {\bf 5}, 423 (1998).
2334: %
2335: \bibitem{alf982}
2336: L.~Alff, A.~Beck, R.~Gross, A.~Marx, S.~Kleefisch, Th.~Bauch, H.~Sato,
2337: M.~Naito, and G.~Koren,
2338: Phys.~Rev.~B {\bf 58}, 11197 (1998).
2339: %
2340: \bibitem{sin98}
2341: S.~Sinha and K.-W.~Ng,
2342: Phys.\,Rev.\,Lett. {\bf 80}, 1296 (1998).
2343: %
2344: \bibitem{wei98}
2345: J.~Y.~T.~Wei, N.-C.~Yeh, D.~F.~Garrigus, and M.~Strasik,
2346: Phys.~Rev.~Lett. {\bf 81}, 2542 (1998).
2347: %
2348: \bibitem{apr99}
2349: M.~Aprili, E.~Badica, and L.~H.~Green,
2350: Phys.~Rev.~Lett. {\bf 83}, 4630 (1999).
2351: %
2352: \bibitem{deutsch99}
2353: R.~Krupke and G.~Deutscher,
2354: Phys.~Rev.~Lett. {\bf 83}, 4634 (1999).
2355: %
2356: \bibitem{Ting110} J.-X. Zhu, B. Friedman, and C.S. Ting, Phys.
2357: Rev. B 59, 3353 (1999).
2358: %
2359: \bibitem{cov00}
2360: M.~Covington and L.~H.~Green,
2361: Phys.~Rev.~B {\bf 62}, 12440 (2000).
2362: %
2363: \bibitem{pairor02}
2364: P.~Pairor and M.~B.~Walker,
2365: Phys.~Rev.~B {\bf 65}, 064507 (2002).
2366: %
2367: \bibitem{greene02}
2368: H.~Aubin, L.H.~Greene, Sha Jian and D.G.~Hinks,
2369: Phys.~Rev.~Lett. {\bf 89}, 177001 (2002).
2370: %
2371: \bibitem{wumou03}
2372: Shin-Tza Wu and Chung-Yu Mou,
2373: Phys. Rev.~B {\bf 67}, 024503 (2003).
2374: %
2375: \bibitem{greene03}
2376: L.~H.~Green, P.~Hentges, H.~Aubin, M.~Aprili, E.~Badica,
2377: M.~Covington, M.~M.~Pafford, G.~Westwood, W.~G.~Klemperer,
2378: S.~Jian, D.~G.~Hinks,
2379: Physica C {\bf 387}, 162 (2003).
2380: %
2381: \bibitem{tan96}
2382: Y.~Tanaka and S.~Kashiwaya,
2383: Phys.~Rev.~B {\bf 53}, 11957 (1996).
2384: %
2385: \bibitem{bbr96}
2386: Yu.~S.~Barash, H.~Burkhardt, and D.~Rainer,
2387: Phys.~Rev.~Lett. {\bf 77}, 4070 (1996).
2388: %
2389: \bibitem{ilichev01}
2390: E. Il'ichev et al.,
2391: Phys.~Rev.~Lett. {\bf 86}, 5369 (2001).
2392: %
2393: \bibitem{blamire03}
2394: G.~Testa, A.~Monaco, E.~Esposito, E.~Sarnelli, D.-J.~Kang, E.~J.~Tarte,
2395: S.~H.~Mennema, and M.~G.~Blamire,
2396: cond-mat/0310727
2397: %
2398: \bibitem{walter98}
2399: H.~Walter, W.~Prusseit, R.~Semerad, H.~Kinder, W.~Assmann, H.~Huber,
2400: H.~Burkhardt, D.~Rainer, and J.~A.~Sauls,
2401: Phys.~Rev.~Lett. {\bf 80}, 3598 (1998).
2402: %
2403: \bibitem{bkk00}
2404: Yu.~S.~Barash, M.~S.~Kalenkov and J.~Kurkij\"arvi,
2405: Phys.~Rev.~B {\bf 62}, 6665 (2000).
2406: %
2407: \bibitem{carr01}
2408: A.~Carrington, F.~Manzano, R.~Prozorov, R.~W.~Giannetta, N.~Kameda,
2409: and T.~Tamegai,
2410: Phys.~Rev.~Lett. {\bf 86}, 1074 (2001).
2411: %
2412: \bibitem{tan00}
2413: S.~Kashiwaya and Y.~Tanaka,
2414: Rep.~Prog.~Phys. {\bf 63}, 1641 (2000).
2415: %
2416: \bibitem{wendin01}
2417: T.~L\"ofwander, V.~S.~Shumeiko and G.~Wendin,
2418: Supercond.~Sci.~Technol. {\bf 14}, R53 (2001).
2419: %
2420: \bibitem{giessibl}F.J. Giessibl, Rev. Mod. Phys. 75, 949 (2003).
2421: %
2422: \bibitem{nagai}
2423: Y.~Nagato and K.~Nagai, Phys. Rev. B {\bf 51}, 16254 (1995).
2424: %
2425: \bibitem{Walker}M. B. Walker and P. Pairor, Phys. Rev. B {\bf 59}, 1421
2426: (1999).
2427: %
2428: \bibitem{Joglekar} Y.N. Joglekar, A.H. Castro Neto, and A.V.
2429: Balatsky, Phys. Rev. Lett. {\bf 92}, 037004 (2004).
2430: %
2431: \bibitem{MorrDemler} D. K. Morr and E. Demler, cond-mat/0010460
2432: %
2433: \bibitem{Tamm}I. Tamm, Phys. Z. Sowjetunion 1, 733 (1932).
2434: %
2435: \bibitem{ginzburg} V.L. Ginzburg, D.A. Kirzhnits, Zh.~\'Eksp. Teor. Fiz. {\bf
2436: 46}, 397 (1964) [Sov.~Phys.~JETP {\bf 19}, 269 (1964)].
2437: %
2438: \bibitem{gantmakher}
2439: V.F. Gantmakher, Y.B. Levinson, {\it Carrier Scattering in Metals and
2440: Semiconductors (Modern Problems in Condensed Matter Sciences, vol. 19)},
2441: Elsevier Science Ltd., 1987.
2442: %
2443: \bibitem{ebz} In the presence of half-filling, surfaces with $(110)$
2444: orientations represent a special case for the square lattice, since
2445: the parallel component of the Fermi momentum coincides in this case with
2446: the edge of the surface-adapted Brillouin zone. In the problems we study,
2447: the role of the edge of the Brillouin zone becomes quite important, in
2448: particular, for this reason. In a more general case one can expect, that
2449: extrema of the quasiparticle spectra would arise at other values of the
2450: momentum component parallel to the surface, which are comparable with
2451: the scale of the Brillouin zone.
2452: %
2453: \bibitem{fominov}
2454: Ya.~V.~Fominov and A.~A.~Golubov, cond-mat/0402381.
2455: %
2456: \bibitem{kos01} S. Kos, Phys. Rev. B {\bf 63}, 214506 (2001).
2457: %
2458: \bibitem{sigrist} C.~Honerkamp, K.~Wakabayashi and M.~Sigrist,
2459: Europhys. Lett. {\bf 50}, 368 (2000).
2460: %
2461: \bibitem{ohashi99} Y. Ohashi, Phys. Rev. B {\bf 60}, 15388 (1999).
2462: %
2463: \bibitem{vanHarlingen02} W.K.~Neils and D.J.~Van~Harlingen,
2464: Phys.~Rev.~Lett. {\bf 88}, 047001 (2002).
2465: %
2466: \bibitem{deutsch02} A.~Sharoni, O.~Millo, A.~Koren, Y.~Dagan,
2467: R.~Beck, G.~Deutscher, G.~Koren, Phys.~Rev.~B {\bf 65}, 134526 (2002).
2468: %
2469: \bibitem{deutsch03} A.~Kohen, G.~Leibovitch, and G.~Deutscher,
2470: Phys.~Rev.~Lett. {\bf 90}, 207005 (2003).
2471: %
2472: \bibitem{tripletfootnote} In terms of standard representation of singlet and triplet pairings
2473: $\hat{\Delta}_s=\Delta i\hat{\sigma}_2$, $\hat{\Delta}_t=(\bm{d}\cdot\hat{\bm{\sigma}})i\hat{\sigma}_2$
2474: one finds for the order parameters we introduced $\Delta_1=\Delta+d_z$, $\Delta_2=\Delta-d_z$,\, $d_x=d_y=0$.
2475: Hence, in the purely singlet case $\Delta_1=\Delta_2$, while for the triplet pairing $\Delta_1=-\Delta_2$.
2476: Otherwise there is a mixture of singlet and triplet pairings.
2477: %
2478: \bibitem{CampuzanoTbreaking} A. Kaminski, S. Rosenknanz, H.M. Fretwell, J.C. Campuzano, Z. Li, H. Raffy,
2479: W.G.  Cullen, H. You, C.G.  Olson, and C.M. Varma, Nature {\bf 416}, 610
2480: (2002).
2481: %
2482: \end{thebibliography}
2483: \end{document}
2484: