cond-mat0603370/tb.tex
1: % D. A. Rowlands, University of Bristol, January 2006
2: % Submitted to J. Phys.:Condens. Matter
3: 
4: \documentclass{iopart}
5: 
6: \usepackage[dvips]{graphics}
7: \usepackage{graphicx}
8: \usepackage{psfrag}
9: 
10: \renewcommand{\u}{\underline}
11: \renewcommand{\b}{\mathbf}
12: \renewcommand{\o}{\overline}
13: \renewcommand{\hat}{\widehat}
14: 
15: \begin{document}
16: 
17: \title{Investigation of the nonlocal coherent-potential approximation}
18: 
19: \author{D.~A.~Rowlands}
20: \address{H.H.~Wills Physics Laboratory, University of Bristol, Bristol BS8 1TL, U.K.}
21: 
22: \begin{abstract}
23: 
24: Recently the nonlocal coherent-potential approximation (NLCPA) has been introduced by Jarrell and Krishnamurthy for describing the electronic
25: structure of substitutionally-disordered systems. The NLCPA provides systematic corrections to the widely used coherent-potential
26: approximation (CPA) whilst preserving the full symmetry of the underlying lattice. Here an analytical and systematic numerical study of the
27: NLCPA is presented for a one-dimensional tight-binding model Hamiltonian, and comparisons with the embedded cluster method (ECM) and
28: molecular coherent potential approximation (MCPA) are made.
29: 
30: \end{abstract}
31: 
32: \pacs{71.23-k, 71.15-m} 
33: 
34: \submitto{\JPCM}
35: 
36: \section{Introduction}
37: 
38: For many years the coherent-potential approximation (CPA) \cite{Soven1} has been widely used for describing the electronic structure of
39: substitutionally-disordered systems. However as a single-site mean-field theory, the CPA is not a fully satisfactory theory of disorder and it
40: leaves much important physics out of consideration. In the CPA a site only feels the average effect of its environment and so, for example,
41: statistical fluctuations in the chemical environment of a site responsible for sharp structure in the density of states (DOS) are not
42: described \cite{Gonis1}. Furthermore the CPA is incapable of describing the effects of chemical short-range order (SRO) and other important
43: environmental effects such as lattice displacements. However a new theory, the nonlocal coherent-potential approximation (NLCPA)
44: \cite{Jarrell1}, has recently been introduced by Jarrell and Krishnamurthy as a way of addressing such problems. The purpose of this paper is
45: to investigate the general validity of the NLCPA by means of numerical calculations with a simple one-dimensional (1D) tight-binding model.
46: 
47: There have been many previous attempts to generalize the CPA, though none have been completely satisfactory \cite{Gonis1}. Of these theories,
48: only the embedded cluster method (ECM) \cite{Gonis6} and the molecular coherent-potential approximation (MCPA) \cite{Tsukada1} will be
49: considered here. The ECM is a method for embedding a cluster of real impurity sites into a medium such at that provided by the CPA. Averaging
50: over an ensemble of the various cluster configurations can therefore give a non self-consistent description of local environment effects. The
51: MCPA goes a step further by enforcing self-consistency with respect to the cluster, and so yields a new effective medium. However this medium
52: has the periodicity of the cluster and so possesses the unsatisfactory property of breaking the symmetry of the underlying lattice. Such
53: supercell periodicity also means the MCPA is computationally prohibitive since it involves a Brillouin zone (BZ) integration which scales
54: with the cluster size, thus preventing its application to realistic systems.
55: 
56: However the recently proposed NLCPA \cite{Jarrell1} appears to possess all the attributes required for a satisfactory cluster generalization
57: of the CPA. It was introduced in the context of a two-dimensional tight-binding model Hamiltonian as the static version of the dynamical
58: cluster approximation (DCA) \cite{Hettler1,Hettler2} used for describing short-range correlations in strongly-correlated electron systems. A
59: recent application has been to describe impurity bound states in disordered d-wave superconductors \cite{Moradian2}. The NLCPA again
60: determines an effective medium via the self-consistent embedding of a cluster, however the problem of maintaining translational invariance is
61: solved by imposing Born-von Karman boundary conditions on the cluster, leading to an effective medium which has the site-to-site
62: translational invariance of the underlying lattice. A further consequence is that the computational difficulties of the MCPA are alleviated
63: since the BZ integration in the NLCPA does not scale as the cluster size increases. 
64: 
65: The NLCPA was subsequently derived by Rowlands \cite{Rowlands1} within the first-principles Korringa-Kohn-Rostoker (KKR) 
66: \cite{Korringa1,Kohn2} multiple scattering formalism, and the resulting KKR-NLCPA \cite{Rowlands1,Rowlands2} method was first implemented for
67: a realistic three-dimensional system ($bcc$ $Cu_{50}Zn_{50}$) in \cite{Rowlands3}. Moreover, the KKR-NLCPA method has recently been combined 
68: with density functional theory, enabling first-principles total energy calculations to be carried out as a function of SRO using 
69: self-consistently determined potentials \cite{Rowlands5}. Such self-consistent-field (SCF)-KKR-NLCPA calculations \cite{Rowlands5} naturally 
70: include a description of the Madelung contribution to the total energy which is missing in the conventional single-site SCF-KKR-CPA, and 
71: important future developments include application to systems with spin \cite{Staunton2}, strain \cite{Gyorffy9}, and valency \cite{Luders1} 
72: fluctuations.
73: 
74: Due to the continued conceptual development of the NLCPA method and its first-principles application to realistic systems in particular, it
75: is important to establish its general validity by addressing the following questions. Firstly, does the NLCPA produce physically meaningful
76: results, in other words are the fluctuations described real? Secondly, does the DOS calculated using the NLCPA converge to the exact result
77: in practice as the cluster size increases, and does it do so systematically? To date, these important questions have not been addressed.
78: However this can be done by carrying out a systematic study of the NLCPA for a simple 1D tight-binding model with diagonal disorder and
79: nearest neighbour hopping. There are three reasons for this. The first is that in 1D the exact result can be obtained against which
80: meaningful comparisons can be made. Secondly, fluctuations are much more significant in 1D and so detailed structure is expected in the DOS
81: which can be accurately interpreted. Thirdly, such a model is computationally very simple and so large clusters can be considered, thus
82: enabling the convergence properties of the theory to be investigated. Such an analysis is carried out in this paper, and any insight gained
83: which could aid the interpretation of realistic calculations is presented.
84: 
85: The outline of this paper is as follows. For completeness and for the purpose of introducing notation, the concept of the effective medium is
86: described first in section~\ref{definitions}. Next, brief derivations of the CPA, ECM, and MCPA are given in sections~\ref{cpa}, \ref{ecm},
87: and \ref{mcpa} respectively. In particular, the equivalence of the `cavity' \cite{Jarrell1,Moradian2} and the `renormalized interactor'
88: \cite{Gonis1} formalism is established. An understanding of these theories is crucial in order to appreciate and assess the advantages and
89: possible limitations of the NLCPA. In section \ref{nlcpa}, a derivation of the NLCPA is given with an attempt to provide a physical picture
90: of the theory, and comparisons with the formalism of sections~\ref{cpa}, \ref{ecm}, and \ref{mcpa} are made where necessary. In
91: section~\ref{graphs}, a numerical investigation of the NLCPA is presented, and results are compared with those of the CPA, ECM, and MCPA.
92: Finally, conclusions are drawn in section~\ref{conclude}.
93: 
94: %--------------------------------------------------------------------------------------------------------------------------------------------
95: \section{Formalism}\label{definitions}
96: 
97: In a basis labelled by the sites of the direct lattice, consider a tight-binding Hamiltonian with matrix elements
98: \begin{equation}\label{hbasis}
99:  H^{ij}=\epsilon^{i}\delta_{ij}+W^{ij}(1-\delta_{ij})
100: \end{equation}
101: and corresponding Dyson equation
102: \begin{equation}\label{propagator}
103:  G^{ij}=G^{ij}_{0}+\sum_{k}G^{ik}_{0}\epsilon^{k}G^{kj}
104: \end{equation} 
105: where $\epsilon^{i}$ is the on-site energy at site $i$, and $W^{ij}$ is the parameter describing the hopping between site $i$ and site $j$. 
106: For the case of a random substitutionally-disordered binary alloy with only diagonal disorder, the set of $\{\epsilon^i\}$ vary from site to
107: site in a random fashion, taking the value $e_A$ with probability $c$ or $e_B$ with probability $1-c$. Denoting a specific configuration of
108: $\{\epsilon^i\}$ by $\gamma$, and the corresponding Green's function for that configuration by $G_{\gamma}$, we are interested in
109: $\left<G_{\gamma}\right>$, the average over all possible configurations. To do that we iterate (\ref{propagator}) in a Born series, average
110: term by term, and resum in the form
111: \begin{equation}\label{dyson}
112:  \left<G^{ij}_{\gamma}\right>=G^{ij}_{0}+\sum_{k,l}G^{ik}_{0}\Sigma^{kl}\left<G^{lj}_{\gamma}\right>
113: \end{equation}
114: where $\Sigma^{ij}$ is the exact \emph{self-energy}.~\footnote{Clearly $\Sigma^{ij}$ is not diagonal in the site index otherwise  we would be
115: making a mean-field approximation by writing the average of a product $\left<\epsilon^{i}G^{ij}_{\gamma}\right>$ in terms of the product of
116: the averages $\left<\epsilon^{i}\right>\left<G^{ij}_{\gamma}\right>$.} Since (\ref{dyson}) is also a Dyson equation, it may be interpreted as
117: describing an \emph{effective medium} in which the site-diagonal part of $\Sigma^{ij}$ decribes an effective on-site energy, and the site
118: off-diagonal part decribes an effective correction to the hopping. The quantities $\left<G^{ij}_{\gamma}\right>$ and $\Sigma^{ij}$ both
119: possess the full translational (and point-group) symmetry of the underlying lattice, and so (\ref{dyson}) may be expressed in reciprocal 
120: space as
121: \begin{equation}\label{dysonk}
122:  \left<G(\b{k})\right>=G_0(\b{k})+G_0(\b{k})\Sigma(\b{k})\left<G(\b{k})\right>
123: \end{equation}
124: The matrix elements of the Green's function are therefore given by the BZ integral
125: \begin{equation}\label{dysonij}
126:  \left<G^{ij}_{\gamma}\right>=\frac{1}{\Omega_{BZ}}\int_{\Omega_{BZ}}d\b{k}\,\left((G_0(\b{k}))^{-1}-\Sigma(\b{k})\right)^{-1}
127:                                 e^{i\b{k}(\b{R}_i-\b{R}_j)}
128: \end{equation}
129: with $(G_0(\b{k}))^{-1}=E-W(\b{k})$. Since it is not feasible to solve the problem exactly, the idea of an effective medium theory is to 
130: determine the best possible approximation to $\Sigma^{ij}$ and $\left<G^{ij}_{\gamma}\right>$.
131: 
132: %--------------------------------------------------------------------------------------------------------------------------------------------
133: \subsection{Coherent-potential approximation}\label{cpa}
134: 
135: The main approximation made by the CPA \cite{Soven1} is to assume a site-diagonal translationally-invariant self-energy 
136: $\o{\epsilon}^{i}\delta_{ij}$, where $\o{\epsilon}^i=\o{\epsilon}$ for all $i$. The CPA effective medium is then described by the equation
137: \begin{equation}\label{dysoncpa}
138:   \o{G}^{ij}=G^{ij}_{0}+\sum_{k}G^{ik}_{0}{\o{\epsilon}}^{k}\o{G}^{kj}
139: \end{equation}
140: In order to determine the medium, let us consider any site $i$. By removing the sum over all sites $k$ and making up for the neglected terms
141: by replacing the free particle Green's function with the cavity Green's function ${\cal{G}}^{ii}$, the site-diagonal part of (\ref{dysoncpa})
142: at site $i$ can be formally rewritten in the form
143: \begin{equation}\label{dysoncavity}
144: \o{G}^{ii}={\cal{G}}^{ii}+{\cal{G}}^{ii}{\o{\epsilon}}^{i}\o{G}^{ii}
145:           =\left(\left({\cal{G}}^{ii}\right)^{-1}-\o{\epsilon}^{i}\right)^{-1}
146: \end{equation}
147: In fact ${\cal{G}}^{ii}$ is related to a quantity known as the renormalized interactor $\Delta^{ii}$ \cite{Gonis1} by
148: \begin{equation}\label{cavdelta}
149:    {\cal{G}}^{ii}=(E-\Delta^{ii})^{-1}
150: \end{equation}
151: where $\Delta^{ii}$ is given by the locator expansion \cite{Gonis1}
152: \numparts
153: \begin{eqnarray}
154: 	\Delta^{ii}&=&\sum_{k{\neq}i}W^{ik}g^{k}W^{ki}+\sum_{k{\neq}i,l{\neq}i}W^{ik}g^{k}W^{kl}g^{l}W^{li}+\cdots\label{delta}\\
155: 	 g^{i}&=&(E-\o{\epsilon}^i)^{-1}\label{gi}	 
156: \end{eqnarray}
157: \endnumparts
158: It can be seen that $\Delta^{ii}$ and therefore ${\cal{G}}^{ii}$ depends only on the medium surrounding site $i$ and is independent of the
159: chemical occupation of $i$ itself. It is therefore straightforward to define the Green's function for a real \emph{impurity} embedded 
160: in the medium simply by replacing the effective site energy $\o{\epsilon}^i$ with a real site energy $\epsilon_{\alpha}^{i}$ at site $i$,
161: where $\alpha=$ $A$ or $B$. From (\ref{dysoncavity}) this is given by
162: \begin{equation}
163: 	G^{ii}_{\alpha}=\left(\left({\cal{G}}^{ii}\right)^{-1}-\epsilon^{i}_{\alpha}\right)^{-1}								 
164: 	               =\left(\o{\epsilon}^{i}-\epsilon^{i}_{\alpha}+(\o{G}^{ii})^{-1}\right)^{-1}
165: \end{equation}
166: Now the CPA demands that
167: \begin{equation}\label{sccpa}
168: 	\sum_{\alpha}P_{\alpha}G^{ii}_{\alpha}=\o{G}^{ii}
169: \end{equation}
170: where $P_{\alpha}$ is the probability that site $i$ is of chemical type $\alpha$. In other words, the replacement of an effective site energy
171: by a real site energy should, on the average, produce no change to the CPA medium. Since the medium is translationally-invariant, it follows
172: from (\ref{dysoncpa}) that $\o{G}^{ii}$ must also be given by the BZ integral
173: \begin{equation}\label{sccpa2}
174:  \o{G}^{ii}=\frac{1}{\Omega_{BZ}}\int_{\Omega_{BZ}}d\b{k}\,\left(E-\o{\epsilon}-W(\b{k})\right)^{-1}
175: \end{equation}
176: The CPA medium is therefore determined from a self-consistent solution of (\ref{sccpa},\ref{sccpa2}).
177: 
178: %--------------------------------------------------------------------------------------------------------------------------------------------
179: \subsection{Embedded cluster method}\label{ecm}
180: 
181: The ECM \cite{Gonis6} is a method for embedding a cluster of real impurity sites in a medium such as that provided by the CPA. First consider
182: (\ref{dysoncpa}) describing the CPA medium, but with the sites $i,j$ lying within some cluster $C$. By restricting the sum over all
183: sites $k$ to run over the sites belonging to the cluster only, and making up for the neglected terms by replacing the free particle Green's
184: function with the cavity Green's function ${\cal{G}}$, (\ref{dysoncpa}) can be formally rewritten in the form
185: \begin{equation}\label{dysonecmIJ}
186: 	\o{G}^{IJ}={\cal{G}}^{IJ}+\sum_{K}{\cal{G}}^{IK}\o{\epsilon}^{K}\o{G}^{KJ}
187: \end{equation}
188: where notation has been introduced so that all sites belonging to the cluster have been denoted by capital letters. In matrix form
189: (\ref{dysonecmIJ}) can be expressed as
190: \begin{equation}\label{dysonecmC}
191: \o{\u{G}}^{CC}={\cal{\u{G}}}^{CC}+{\cal{\u{G}}}^{CC}\o{\u{\epsilon}}^{C}\o{\u{G}}^{CC}
192:               =\left(({\cal{\u{G}}}^{CC})^{-1}-\o{\u{\epsilon}}^{C}\right)^{-1}
193: \end{equation}
194: where all matrices are in the space of the sites belonging to the cluster. In analogy to (\ref{cavdelta}) in the derivation of the
195: CPA, ${\cal{\u{G}}}^{CC}$ is related to the \emph{cluster} renormalized interactor $\u{\Delta}^{CC}$ \cite{Gonis1} by the 
196: expression
197: \begin{equation}
198:    {\cal{\u{G}}}^{CC}=(\u{E}-\u{\Delta}^{CC})^{-1}
199: \end{equation}
200: where $\Delta^{IJ}$, the matrix elements of $\u{\Delta}^{CC}$, are given by the expansion
201: \begin{equation}
202: \Delta^{IJ}=\sum_{k\not{\in}C}W^{Ik}g^{k}W^{kJ}+\sum_{k\not{\in}C,l\not{\in}C}W^{Ik}g^{k}W^{kl}g^{l}W^{lJ}+\cdots 
203: \end{equation}
204: and $g^i$ is given by (\ref{gi}). It can be seen that $\u{\Delta}^{CC}$ and therefore ${\cal{\u{G}}}^{CC}$ depends only on the
205: medium surrounding the cluster $C$ and is independent of the chemical occupation of $C$ itself. The Green's function for an \emph{impurity 
206: cluster} embedded in the CPA medium can now be straightforwardly defined simply by replacing the effective CPA cluster-site energy matrix 
207: $\o{\u{\epsilon}}^C$ with a matrix $\u{\epsilon}^{C}_{\gamma}$ containing a configuration $\gamma$ of real site energies 
208: $\{\epsilon_{\alpha}^{I}\}$ at each cluster site $I$, where $\alpha=$ $A$ or $B$. From (\ref{dysonecmC}) this is given by
209: \begin{equation}\label{ecmimpC}
210: 	\u{G}^{CC}_{\gamma}=\left(\left({\cal{\u{G}}}^{CC}\right)^{-1}-\u{\epsilon}^{C}_{\gamma}\right)^{-1}
211: 	                   =\left(\o{\u{\epsilon}}^{C}-\u{\epsilon}^{C}_{\gamma}+(\o{\u{G}}^{CC})^{-1}\right)^{-1}
212: \end{equation}
213: Once the CPA medium is known, the Green's function for an impurity cluster embedded in the medium can be straightforwardly determined through
214: (\ref{ecmimpC}), and properties can then be calculated at any desired cluster site. However an average over all possible cluster
215: configurations $\gamma$ is usually taken, in which case it will be found that $\left<\u{G}^{CC}_{\gamma}\right>\neq{\o{\u{G}}^{CC}}$ for a
216: cluster size $N_c>1$. This is due to an average effect on the electron hopping within the cluster caused by the disorder configurations. Thus
217: $\o{\u{\epsilon}}^{C}$ gains an off-diagonal correction and could be interpreted as a (non-self-consistently determined) \emph{cluster
218: self-energy} $\u{\Sigma}^{C}$ with matrix elements $\Sigma^{IJ}$. However this is defined only at the cluster sites, with the single-site CPA
219: self-energy $\o{\epsilon}^{i}$ situated at all other sites, and hence there is no new effective medium. To do that, self-consistency with
220: respect to the cluster must be enforced. One way forward is the MCPA \cite{Tsukada1}, as described in the next subsection.
221: 
222: %--------------------------------------------------------------------------------------------------------------------------------------------
223: \subsection{Molecular coherent-potential approximation}\label{mcpa}
224: 
225: As described above, a cluster self-energy may be defined by averaging over an ensemble of impurity cluster configurations embedded in a
226: medium. The idea of the MCPA \cite{Tsukada1} is to determine an effective medium with a periodically-repeating cluster self-energy. Such a
227: medium may be defined by first dividing the lattice up into a collection of identical non-overlapping clusters, and then writing the Green's
228: function in the form
229: \begin{equation}\label{dysonmcpa}
230:     \o{\u{G}}^{CC'}=\u{G}_{0}^{CC'}+\sum_{C''}\u{G}_{0}^{CC''}\left(\u{\Sigma}^{C''}+\u{W}^{C''C''}\right)\o{\u{G}}^{C''C'}
231: \end{equation}
232: These cluster matrices have site matrix elements
233: \begin{equation}
234:   \left[\u{\o{G}}^{CC'}\right]_{ij}=\o{G}^{ij} , \quad 
235: 	\left[\u{G}_{0}^{CC'}\right]_{ij}=G_{0}^{ij} , \quad i\in C,\ j\in C' 
236: \end{equation}
237: \begin{equation}
238: 		\left[\u{\Sigma}^{C}\delta_{CC'}\right]_{ij}=\Sigma^{ij} , \quad
239: 		\left[\u{W}^{CC}\delta_{CC'}\right]_{ij}=W^{ij} , \quad i,j \in C 
240: \end{equation}
241: The main approximation made here has been to assume that the (as of yet undetermined) cluster self-energy is \emph{cluster-diagonal}. This
242: means there are self-energy terms relating sites in the same cluster, but none relating sites in different clusters. The single-site
243: translational invariance of the underlying lattice is therefore broken. However, the MCPA does require the medium to be invariant upon
244: translation by a cluster so that it has the periodicity of a `super-lattice'. As such, $\u{\Sigma}^{C}$ is independent of the cluster label
245: $C$, and the Green's function $\o{\u{G}}^{CC'}$ depends only on the difference between $C$ and $C'$. The real-space cluster dependence of
246: (\ref{dysonmcpa}) may therefore be removed using the Fourier transform
247: \begin{equation}\label{mcpaft}
248:     \o{\u{G}}^{CC}(\b{q})=\sum_{{C'}}\o{\u{G}}^{CC'}e^{-i\b{q}(\b{R}_{C}-\b{R}_{C'})}
249: \end{equation}
250: where $\b{R}_{C}-\b{R}_{C'}$ is the vector distance between the centres of clusters $C$ and $C'$, and $\b{q}$ is a vector in the (smaller) BZ
251: of the super-lattice. Equation~(\ref{dysonmcpa}) now becomes
252: \begin{equation}\label{dysonq}
253:     \o{\u{G}}^{CC}(\b{q})=\u{G}_{0}^{CC}(\b{q})+\u{G}_{0}^{CC}(\b{q})\left(\u{\Sigma}^{C}+\u{W}^{CC}\right)\o{\u{G}}^{CC}(\b{q})
254: \end{equation}
255: where all matrices have the dimension of the cluster size only, and the inter-cluster dependence is expressed through the reciprocal-space 
256: vector $\b{q}$. 
257: 
258: The next step is to embed an ensemble of impurity cluster configurations into the medium defined above. This may done using the ECM of 
259: section~\ref{ecm}, yielding the impurity cluster Green's function
260: \begin{equation}
261: 	\u{G}^{CC}_{\gamma}=\left(\u{\Sigma}^{C}-\u{\epsilon}^{C}_{\gamma}+(\o{\u{G}}^{CC})^{-1}\right)^{-1}
262: \end{equation}
263: However, there are two main differences. Firstly, the cluster self-energy $\u{\Sigma}^C$ replaces the effective CPA cluster-site energy
264: matrix $\u{\epsilon}^C$. Secondly, the medium outside the cluster comprises of a periodically repeating cluster self-energy rather than
265: single-site CPA effective site energies, and hence $\u{\Delta}^{CC}$ has a slightly different expansion~\cite{Gonis1}. The final step is to
266: generalize the CPA argument and demand that the average of $\u{G}^{CC}_{\gamma}$ over all $\gamma$ be equal to the Green's function for the
267: medium itself i.e.
268: \begin{equation}\label{scmcpa}
269:     \left<\u{G}^{CC}_{\gamma}\right>=\o{\u{G}}^{CC}\quad{or}\quad\sum_{\gamma}P_{\gamma}G^{IJ}_{\gamma}=\o{G}^{IJ}\quad{\forall\,{I,J}\in{C}}
270: \end{equation}
271: For a cluster containing $N_c$ sites, the number of configurations will be $2^{N_c}$ for a binary alloy. Since the medium has the periodicity
272: of the cluster, it follows from (\ref{mcpaft},\ref{dysonq}) that $\o{\u{G}}^{CC}$ must also satisfy
273: \begin{eqnarray}\label{bzmcpa}
274: \o{\u{G}}^{CC}=\frac{1}{\Omega_{BZ}}\int_{\Omega_{BZ}}d\b{q}\left(\u{E}-\u{\Sigma}^{C}-\u{W}^{CC}-\u{W}^{CC}(\b{q})\right)^{-1}
275: \end{eqnarray}
276: where the integral is over the first BZ of the super-lattice. Therefore the MCPA medium is determined from a self-consistent
277: solution of (\ref{scmcpa},\ref{bzmcpa}).
278: 
279: The MCPA appears to be a natural generalization of the single-site CPA, but it does have some major disadvantages. Specifically, the
280: intra-cluster hopping is now different from the inter-cluster hopping expressed through $\u{W}^{CC}(\b{q})$ and so measuring properties at 
281: sites situated at the boundary of the cluster will in general give different results than sites at the centre of the cluster, for example. 
282: The supercell periodicity also causes problems in obtaining a well-defined spectral function, and leads to the computationally-demanding BZ
283: integration in (\ref{bzmcpa}) which scales as $N_c$ increases. 
284: 
285: %--------------------------------------------------------------------------------------------------------------------------------------------
286: \subsection{Nonlocal coherent-potential approximation}\label{nlcpa}
287: 
288: Like the MCPA, the NLCPA \cite{Jarrell1} determines an effective medium via the self-consistent embedding of a cluster. However, the key
289: difference is that \emph{Born-von Karman boundary conditions} are imposed on the cluster, leading to an effective medium which has the
290: translational invariance of the underlying lattice. The first step is to solve the problem of an isolated cluster with Born-von Karman
291: boundary conditions imposed. 
292: 
293: \subsubsection{Cluster with Born-von Karman boundary conditions}\label{bvk}
294: 
295: \begin{figure}[!]
296:  \begin{center}
297:   \psfrag{a}[B1][B1][1.5][0]{$a$}
298:   \psfrag{-pi/a}[B1][B1][1.5][0]{$\frac{-\pi}{a}$}
299: 	\psfrag{pi/a}[B1][B1][1.5][0]{$\frac{\pi}{a}$}
300:  \scalebox{0.6}{\includegraphics{fig1.eps}}   
301:  \caption{(a) Real space 1D tile (denoted by double-headed arrow of length $4a$) for a $N_c=4$ cluster. Sites are denoted by open circles, 
302: 	and $a$ is the lattice constant. (b) Set of cluster momenta (periodic) denoted by closed circles for the $N_c=4$ cluster. The tiles centred
303: 	at the cluster momenta are denoted by arrows and the solid line is the first BZ. The part of the tile centred at $\pi/a$ that lies outside
304: 	the first BZ can be translated by reciprocal lattice vectors into the first BZ to lie between $-\pi/a$ and $-\pi/4a$. (c) Set of cluster 
305: 	momenta (anti-periodic) denoted by closed circles for the $N_c=4$ cluster. Again the tiles centred at the cluster momenta are denoted by 
306: 	arrows and the solid line is the first BZ.}\label{cluster}
307:  \end{center}
308: \end{figure}
309: 
310: Conventionally in solid state theory the problem of a lattice comprising of a large number of sites with Born-von Karman boundary conditions
311: is considered \cite{Ashcroft1}. In 1D, for example, this essentially means a very long chain of sites where one end of the chain maps round
312: to the other end. If the lattice constant is $a$, then the first BZ extends from $-\pi/a$ to $+\pi/a$ in reciprocal space, and the number of
313: $\b{k}$ points in the BZ is equal to the number of sites in the chain. It follows that if the length of the chain is now decreased to contain
314: only say a cluster of $N_c$ sites, then the number of $\b{k}$ points in the BZ will then equal $N_c$. These are referred to as the set of
315: \emph{cluster momenta} $\{\b{K}_n\}$, where $n=1,..,N_c$. However, the boundaries of the BZ remain the same since the lattice constant is
316: unchanged. The real-space cluster sites $\{I\}$ (denoted by capital letters) and the corresponding set of cluster momenta $\{\b{K}_n\}$
317: satisfy the relation \cite{Jarrell1} 
318: \begin{equation}\label{IJK}
319: 	\frac{1}{N_c}\sum_{\b{K}_n}e^{i\b{K}_n(\b{R}_I-\b{R}_J)}=\delta_{IJ}
320: \end{equation}
321: for which the conventional lattice Fourier transform is recovered when $N_c\rightarrow\infty$. For example, in 1D the (periodic) solutions
322: are 
323: \begin{eqnarray}
324: 	\{\b{R}_I\}=&(I-1)a& \ \ \ I=1,..,N_c \nonumber\\
325: 	\{\b{K}_n^{P}\}=&\left(\frac{2n-N_c}{N_c}\right)\frac{\pi}{a}& \ \ \ n=1,..,N_c
326: \end{eqnarray}
327: for $N_c$ even. To date, only periodic solutions have been considered in the literature. However, for a given cluster size there is also a 
328: solution of (\ref{IJK}) corresponding to anti-periodic Born-von Karman boundary conditions (where in 1D, for example, the wavefunction at  
329: one end of the chain maps back to minus the value at the other end). In 1D, the anti-periodic set of cluster momenta are given by
330: \begin{equation}
331: 	\{\b{K}_n^{AP}\}=\left(\frac{2n-N_c-1}{N_c}\right)\frac{\pi}{a} \ \ \ n=1,..,N_c
332: \end{equation}
333: A 1D example for $N_c=4$ is shown in figure~\ref{cluster}. Notice that $\{\b{K}_n^{P}\}$ always includes the origin in reciprocal space,
334: whilst $\{\b{K}_n^{AP}\}$ are shifted (by $-\pi/(N_c\,{a})$ in 1D) and lie symmetric about the origin. For the remainder of this derivation
335: it is however convenient to simply choose one of the set of solutions above. A discussion concerning this choice is given in 
336: section~\ref{solutions}. Cluster quantities that are translationally-invariant can be related in real and reciprocal space through 
337: (\ref{IJK}), for example for the cluster self-energy we have
338: \begin{eqnarray}\label{clusterselfenergy}
339: 	\Sigma_{cl}^{IJ}=\frac{1}{N_c}\sum_{\b{K}_n}\Sigma_{cl}(\b{K}_n)e^{i\b{K}_n(\b{R}_I-\b{R}_J)} \nonumber\\
340: 	\Sigma_{cl}(\b{K}_n)=\sum_{J}\Sigma_{cl}^{IJ}e^{-i\b{K}_n(\b{R}_I-\b{R}_J)}
341: \end{eqnarray}
342: 
343: \subsubsection{Mapping cluster to lattice}
344: 
345: \begin{figure}[!]
346:  \begin{center}
347:  \scalebox{0.6}{\includegraphics{fig2.eps}}   
348:  \caption{(a) Self-energy $\Sigma^{IJ}$ (denoted by double-headed arrows) for the case of a $N_c=2$ cluster in the MCPA, for $I\neq{J}$. The 
349:  vertical lines denote the boundaries of the tiles. (b) Same as (a) but for the NLCPA.}\label{selfenergy}
350:  \end{center}
351: \end{figure}
352: 
353: When mapping the cluster problem to the lattice, there are certain conditions on the geometry of the chosen cluster which must be satisified
354: in order that that the translational and point-group symmetry of the lattice is preserved \cite{Jarrell1}. Specifically, if the cluster sites
355: are now considered as `embedded' in the lattice, then it must be possible to surround them with a tile which has this point-group symmetry
356: and can be periodically repeated to fill out all space. This is analagous to the conventional Wigner-Seitz tile \cite{Ashcroft1} used to
357: surround a single site, and restricts the allowed values of $N_c$ for a given lattice. The construction for choosing appropriate clusters and
358: tiles was first explained in \cite{Jarrell1} for a 2D square lattice and was generalized to realistic 3D lattices in
359: \cite{Rowlands1,Rowlands2,Rowlands3}.~\footnote{This construction would also apply to the MCPA if aiming to preserve point-group symmetry.}
360: In 1D, the tiles are simply lines (see figure~\ref{cluster}).
361: 
362: Next, note that the periodicity of the real-space tiles define a reciprocal lattice, and each reciprocal lattice vector $\b{K}$ will
363: correspondingly be centred at a reciprocal-space tile. For $N_c=1$, this would correspond to the usual tiling of the conventional reciprocal
364: lattice with Brillouin zones. However, for $N_c>1$, reciprocal space will be more densely filled with $\b{K}$ points and the corresponding 
365: tiles will be smaller (by a factor of $N_c$). In the MCPA, such a tile would be the BZ of the `superlattice' (see section~\ref{mcpa}).
366: However, the idea of the NLCPA is to use $N_c$ such tiles to fill out the original BZ of the underlying lattice (see figure~\ref{cluster} for
367: a 1D example), thus preserving translational invariance. In fact, the relevant $\b{K}$ points will be the set of cluster momenta $\{\b{K}_n\}$.
368: 
369: The mapping of the cluster problem to the lattice may now be achieved as follows. In reciprocal space, consider the exact lattice self-energy
370: $\Sigma(\b{k})$ of (\ref{dysonk}). The first step is to average $\Sigma(\b{k})$ over the momenta $\b{q}$ within each of the $N_c$
371: tiles. This results in a `coarse-grained' lattice self-energy $\Sigma(\b{K}_n)$ which has a constant but different value within each tile. On
372: the other hand, the cluster self-energy $\Sigma_{cl}(\b{K}_n)$ is defined only at the cluster momenta $\{\b{K}_n\}$. The main approximation
373: made by the NLCPA is to set $\Sigma(\b{K}_n)$ to be equal to the value $\Sigma_{cl}(\b{K}_n)$ within each tile $n$ i.e.
374: \begin{equation}
375: 	\frac{1}{\Omega_{\b{K}_n}}\int_{\Omega_{\b{K}_n}}d\b{q}\,\Sigma(\b{K}_n+\b{q})=\Sigma(\b{K}_n)\simeq\Sigma_{cl}(\b{K}_n)	
376: \end{equation}
377: In other words, the lattice self-energy is approximated by that obtained from the cluster. The NLCPA approximation to the lattice Green's
378: function in reciprocal space may now be represented by summing over the dispersion within each tile, yielding the set of coarse-grained 
379: values
380: \begin{equation}\label{GK}
381: 	\o{G}(\b{K}_n)=\frac{N_c}{\Omega_{BZ}}\int_{\Omega_{\b{K}_n}}d\b{k}\left(E-W(\b{k})-\Sigma(\b{K}_n)\right)^{-1}
382: \end{equation}
383: which are straightforward to calculate since $\Sigma(\b{K}_n)$ is constant within each tile $\Omega_{\b{K}_n}$. Finally, using (\ref{IJK}), 
384: the real space Green's function at the cluster sites becomes
385: \begin{equation}\label{GIJK}
386: 	\o{G}^{IJ}=\frac{1}{\Omega_{BZ}}\sum_{\b{K}_n}\int_{\Omega_{\b{K}_n}}
387: 	            d\b{k}\left(E-W(\b{k})-\Sigma(\b{K}_n)\right)^{-1}e^{i\b{K}_n(\b{R}_I-\b{R}_J)}
388: \end{equation}
389: The reason for coarse-graining the Green's function is to preserve causality by removing phase factors within the cluster, however these can 
390: be reintroduced after the medium has been determined (see section~\ref{observables}).
391: 
392: A physical picture of the resulting NLCPA medium may be seen by examining the consequences of approximating the exact self-energy
393: $\Sigma(\b{k})$ by $\Sigma(\b{K}_n)$. Firstly, the self-energy is now restricted to act within the \emph{range}~\footnote{The self-energy
394: $\Sigma^{IJ}$ only takes into account exactly nonlocal correlations that are within the range of the cluster size. However, when the medium
395: is determined self-consistently, $\Sigma^{IJ}$ will include the effects of physics on a longer length scale at the mean-field level.} of the
396: cluster sites $\{I,J\}$ only (a more precise correlation length can be defined using Nyquist's sampling theorem \cite{Elliot1,Jarrell1}).
397: However, an electron can now propagate to any site in the lattice via the dispersion $W(\b{k})$, and will experience an identical self-energy
398: $\Sigma^{IJ}$ at that site. Thus $\Sigma^{IJ}$ remains a translationally-invariant quantity which depends only on the distance between sites
399: $I$ and $J$, now within the range of the cluster size, but independent of which site in the lattice is chosen to be site $I$. ($\Sigma^{IJ}$
400: will also possess the point-group symmetry of the underlying lattice provided that the clusters and tiles have been chosen correctly). An
401: illustration of this for the simplest case of a $N_c=2$ cluster in one dimension is shown in figure~\ref{selfenergy}.
402: 
403: \subsubsection{Impurity problem}
404: 
405: To determine the medium, the impurity problem must be solved. The first step is to define the reciprocal space cavity Green's function 
406: ${\cal{G}}(\b{K}_n)$ via the Dyson equation
407: \begin{equation}\label{dysonKn}
408:     \o{G}(\b{K}_n)={\cal{G}}(\b{K}_n)+{\cal{G}}(\b{K}_n)\Sigma(\b{K}_n)\o{G}(\b{K}_n)
409: \end{equation}
410: In diagrammatic terms ${\cal{G}}(\b{K}_n)$ is introduced to `avoid over-counting self-energy diagrams on the cluster' \cite{Hettler1}.
411: Equation~(\ref{dysonKn}) can of course also be expressed in real space by applying the Fourier transform (\ref{IJK}) to yield
412: \begin{equation}\label{dysonreal}
413:     \o{G}^{IJ}={\cal{G}}^{IJ}+\sum_{K,L}{\cal{G}}^{IK}\Sigma^{KL}\o{G}^{LJ}
414: \end{equation}
415: where ${\cal{G}}^{IJ}$ is the real space cavity Green's function. In a similar manner to that introduced in the ECM and MCPA, here 
416: ${\cal{G}}^{IJ}$ is independent of the chemical occupation of the cluster itself. However, the important difference is that it is not
417: possible to give an explicit real space expansion for ${\cal{G}}^{IJ}$ in terms of the cluster renormalized interactor as was possible in the
418: ECM and MCPA. The reason for this is due to the translational invariance of the self-energy in the NLCPA, and is an example of the so-called
419: `embedding problem' \cite{Gonis5}. If the ECM were used here to accomodate an impurity cluster in the NLCPA medium, then self-energy terms
420: linking the cluster sites to the medium sites would have to be broken (see figure~\ref{selfenergy}). Such a situation does not occur in the
421: MCPA since the self-energy is cluster diagonal and hence there are no self-energy terms linking sites in a cluster to the rest of the medium
422: in that case. Although the NLCPA cavity Green's function ${\cal{G}}^{IJ}$ is a mathematically well-defined quantity, it is not the same
423: quantity as that introduced in the ECM. The embedding of a cluster with Born-von Karman boundary conditions in the NLCPA medium should
424: therefore be viewed as a mathematical construction rather than a true embedding in the conventional sense. Nevertheless, the NLCPA impurity
425: cluster Green's function may be defined in analogy to (\ref{ecmimpC}) by replacing the cluster self-energy in (\ref{dysonreal}) with a 
426: particular configuration of site energies $\{\epsilon^I_{\alpha}\}$ i.e.
427: \begin{equation}
428:     G^{IJ}_{\gamma}={\cal{G}}^{IJ}+\sum_{K}{\cal{G}}^{IK}\epsilon_{\alpha}^{K}G^{KJ}_{\gamma}
429: \end{equation}
430: The NLCPA self-consistency condition is equivalent to that of the MCPA given in (\ref{scmcpa}) i.e.
431: \begin{equation}\label{scnlcpa}
432:     \sum_{\gamma}P_{\gamma}G^{IJ}_{\gamma}=\o{G}^{IJ}
433: \end{equation}
434: where $P_{\gamma}$ is the probability of configuration $\gamma$ occuring. The effective medium is therefore determined from a self-consistent
435: solution of (\ref{GIJK},\ref{scnlcpa}). The cluster probabilities depend on the SRO parameter $\alpha$ and so SRO may be included by
436: appropriately weighting the configurations in (\ref{scnlcpa}), provided that translational-invariance is preserved. Finally note that the
437: NLCPA formalism reduces to the CPA for $N_c=1$ and becomes exact as $N_c\rightarrow\infty$ since this would amount to solving the exact
438: problem described by (\ref{dyson},\ref{dysonk}).
439: 
440: \subsubsection{Calculation of Observables}\label{observables}
441: 
442: Once the medium has been determined through (\ref{GIJK},\ref{scnlcpa}), there is no longer any need to coarse-grain the Green's 
443: function via (\ref{GK}). Now, the Green's function may be calculated at any point in the BZ through
444: \begin{equation}\label{gkdisc}
445: 	\o{G}(\b{k})=\left(E-W(\b{k})-\Sigma(\b{K}_{n})\right)^{-1}
446: \end{equation}
447: and correspondingly at any sites $i,j$ in the lattice by
448: \begin{equation}\label{gijdisc}
449:   \o{G}^{ij}=\frac{1}{\Omega_{BZ}}\int_{\Omega_{BZ}}d\b{k}\left(E-W(\b{k})-\Sigma(\b{K}_n)\right)^{-1}e^{i\b{k}(\b{R}_i-\b{R}_j)}
450: \end{equation}
451: In (\ref{gkdisc},\ref{gijdisc}) above, $\Sigma(\b{K}_n)$ takes the appropriate value within each tile $n$. The configurationally-averaged DOS
452: per site is given by the usual expression
453: \begin{equation}
454:   \o{n}(E)=-\frac{1}{\pi}\,Im\,\,\o{G}^{II}
455: \end{equation}
456: where $\o{G}^{II}$ may be calculated from (\ref{GIJK}) or from (\ref{gijdisc}) above. However, when calculating site-off diagonal observables
457: such as the spectral function, notice that $\Sigma(\b{K}_n)$ taking the appropriate (constant) value within each tile in (\ref{gkdisc}) leads
458: to discontinuities in $\o{G}(\b{k})$ at the tile boundaries. In order to remedy this, Maier~\etal \cite{Maier3} proposed a scheme
459: (within the DCA) to interpolate the self-energy using maximum entropy, however this can lead to causality being violated when the
460: fluctuations are large. An alternative scheme for removing such discontinuities has been proposed by Batt~\etal \cite{Batt1} (see
461: section~\ref{solutions} below). 
462: 
463: \subsubsection{Periodic and anti-periodic solutions}\label{solutions}
464: 
465: It was mentioned in section~\ref{bvk} that for a given cluster size there are two possible sets of cluster momenta which may be used,
466: $\{\b{K}_n^{P}\}$ and $\{\b{K}_n^{AP}\}$, corresponding to periodic or anti-periodic Born-von Karman boundary conditions respectively. In the
467: literature to date, only periodic boundary conditions have been considered. The validity of calculations using only $\{\b{K}_n^{P}\}$ is
468: analyzed in section~\ref{graphs}. However, it has been suggested that one should perform separate calculations using $\{\b{K}_n^{P}\}$ and
469: $\{\b{K}_n^{AP}\}$, and then average over both results \cite{Jarrell2}. The philosophy behind this argument is that this has the effect of
470: increasing the number of tiles, leading to a better sampling of the BZ. For example, in 1D the number of tiles is effectively doubled, as can
471: be seen from figure~\ref{cluster}. The validity of this suggestion is also examined in section~\ref{graphs}. 
472: 
473: However, Batt~\etal \cite{Batt1} have proposed a more sophisticated scheme for combining both solutions which also removes the
474: discontinuities in $\o{G}(\b{k})$ mentioned in section~\ref{observables} above. Firstly, let us label the Green's function (given by
475: (\ref{gkdisc})) for the periodic and anti-periodic solutions by $\o{G}_{P}(\b{k})$ and $\o{G}_{AP}(\b{k})$ respectively. Next, observe that
476: by construction $\o{G}_{P}(\b{k})$ is a better approximation to the exact result in the region of reciprocal space close to each of the
477: points $\{\b{K}_n^{P}\}$ (in fact $\o{G}_{AP}(\b{k})$ has discontinuities at $\{\b{K}_n^{P}\}$). Similarly $\o{G}_{AP}(\b{k})$ is a better
478: approximation to the exact result close to each of the points $\{\b{K}_n^{AP}\}$. This suggests that, in brief, one should construct a new
479: Green's function which follows $\o{G}_{P}(\b{k})$ close to each of the points $\{\b{K}_n^{P}\}$, and follows $\o{G}_{AP}(\b{k})$ close to
480: each of the points $\{\b{K}_n^{AP}\}$, and interpolates in-between. Such a combined solution is guaranteed to be causal since it will always
481: lie between the two causal extremes $\o{G}_{P}(\b{k})$ and $\o{G}_{AP}(\b{k})$. Furthermore, the problem of discontinuities is automatically
482: removed since neither solution is followed where it has a discontinuity. Full details together with an implementation of this scheme are
483: given in \cite{Batt1}.
484: 
485: %--------------------------------------------------------------------------------------------------------------------------------------------
486: \section{Results}\label{graphs}
487: 
488: \begin{figure}[!]
489:  \begin{center}
490:  \begin{tabular}{cc}
491:  \psfrag{(a)}[B1][B1][2][0]{(a)}\scalebox{0.33}{\includegraphics{fig3a.eps}} &
492:  \psfrag{(c)}[B1][B1][2][0]{(c)}\scalebox{0.33}{\includegraphics{fig3c.eps}} \\
493:  \psfrag{(b)}[B1][B1][2][0]{(b)}\scalebox{0.33}{\includegraphics{fig3b.eps}} & 
494:  \psfrag{(d)}[B1][B1][2][0]{(d)}\scalebox{0.33}{\includegraphics{fig3d.eps}} 
495:  \end{tabular}         
496:  \caption{(a) DOS (as a function of energy) for a pure material comprising of $A$ sites, with $\epsilon_A=+2.0$. 
497:  					(b) DOS for a pure material comprising of $B$ sites, with $\epsilon_B=-2.0$. 
498: 					(c) Exact DOS results for a random $A_{50}B_{50}$ alloy of the pure materials above. 
499: 					(d) DOS for the same $A_{50}B_{50}$ alloy obtained using the CPA.}\label{pureexactcpa}
500:  \end{center}
501: \end{figure}
502: 
503: In this section results are presented for a 1D tight-binding model with diagonal disorder only and nearest-neighbour hopping. Calculations
504: for such a model were first performed with the NLCPA by Moradian \cite{Moradian1} using $N_c=4$ and $N_c=8$ clusters. Here the random site
505: energies can take the values $\epsilon_A=+2.0$ and $\epsilon_B=-2.0$, and the nearest neighbour hopping parameter is taken to be $W=1.0$.
506: Figure~\ref{pureexactcpa}(a) and figure~\ref{pureexactcpa}(b) show the DOS results for a pure material comprising of energies $\epsilon_A$ or
507: $\epsilon_B$ respectively. Notice that the pure bands just touch at $E=0.0$, and so a theory for an alloy of the two pure materials must deal
508: here with this difficult `split-band' regime. Indeed, from the Saxon-Hutner theorem \cite{Ziman1}, the gap at $E=0.0$ for an alloy of the
509: two pure materials should correspondingly be just vanishing. 
510: 
511: \subsection{Exact Result and the CPA}
512: 
513: Figure~\ref{pureexactcpa}(c) shows exact DOS results for a random $A_{50}B_{50}$ alloy of the two pure materials shown in
514: figures~\ref{pureexactcpa}(a),(b) obtained using the negative eigenvalue theorem \cite{Dean1}. The results have been displayed using centred
515: histograms at 400 points from $E=-4.0$ to $E=+4.0$ i.e.~an energy increment of 0.01 (a higher resolution yields too much structure for
516: meaningful comparisons to be made). As expected, the DOS is highly structured and the gap at $E=0.0$ is just vanishing. 
517: 
518: Next, figure~\ref{pureexactcpa}(d) shows a CPA calculation for the same $A_{50}B_{50}$ alloy, displayed using centred histograms at the same
519: energy resolution as the exact result. The CPA result is very smooth due to the neglect of nonlocal fluctuations in the disorder
520: configurations. However, one of the central successes of the CPA is its ability to describe the split-band regime. Indeed the CPA correctly
521: separates the DOS here into two sub-bands, but it is clear that it fails to correctly describe the onset of such split-band behaviour. This
522: is because the contributions to the DOS near band edges are due to large clusters of like atoms which are not described by the CPA
523: \cite{Gonis1}. 
524: 
525: \subsection{Cluster theories}
526: 
527: \begin{figure*}[!]
528:  \begin{center}
529:  \begin{tabular}{ccc}
530:    \psfrag{ECM}[B1][B1][2][0]{ECM}\scalebox{0.33}{\includegraphics{ecm2.eps}}     
531:  & \scalebox{0.33}{\includegraphics{ecm4.eps}}     
532:  & \scalebox{0.33}{\includegraphics{ecm6.eps}}\\
533:    \psfrag{MCPA}[B1][B1][2][0]{MCPA centre}\scalebox{0.33}{\includegraphics{mcpa2.eps}}    
534:  & \scalebox{0.33}{\includegraphics{mcpa4.eps}}   
535:  & \scalebox{0.33}{\includegraphics{mcpa6.eps}}\\ 
536:    \psfrag{MCPA}[B1][B1][2][0]{MCPA average}\scalebox{0.33}{\includegraphics{mcpa2avg.eps}} 
537:  & \scalebox{0.33}{\includegraphics{mcpa4avg.eps}} 
538:  & \scalebox{0.33}{\includegraphics{mcpa6avg.eps}}\\
539:    \psfrag{NLCPA}[B1][B1][2][0]{NLCPA periodic}\scalebox{0.33}{\includegraphics{tb2p.eps}}
540:  & \scalebox{0.33}{\includegraphics{tb4p.eps}}     
541:  & \scalebox{0.33}{\includegraphics{tb6p.eps}}\\
542:    \psfrag{NLCPA}[B1][B1][2][0]{NLCPA anti-periodic}\scalebox{0.33}{\includegraphics{tb2a.eps}} 
543:  & \scalebox{0.33}{\includegraphics{tb4a.eps}}     
544:  & \scalebox{0.33}{\includegraphics{tb6a.eps}}\\ 
545:    \psfrag{Nc2}[B1][B1][2.5][0]{$N_C=2$}\psfrag{NLCPA}[B1][B1][2][0]{NLCPA average}\scalebox{0.33}{\includegraphics{tb2avg.eps}}
546:  & \psfrag{Nc4}[B1][B1][2.5][0]{$N_C=4$}\scalebox{0.33}{\includegraphics{tb4avg.eps}}   
547:  & \psfrag{Nc6}[B1][B1][2.5][0]{$N_C=6$}\scalebox{0.33}{\includegraphics{tb6avg.eps}}    
548:  \end{tabular}         
549:  \caption{Configurationally-averaged DOS per site as a function of energy (in units of the bandwidth) for the various cluster theories (top 
550:  to bottom) with cluster sizes $N_c=2,4,6$ (left to right).}\label{cluster246}
551:  \end{center}
552: \end{figure*}
553: 
554: \begin{figure*}[!]
555:  \begin{center}
556:  \begin{tabular}{ccc}
557:    \psfrag{ECM}[B1][B1][2][0]{ECM}\scalebox{0.33}{\includegraphics{ecm8.eps}}     
558:  & \scalebox{0.33}{\includegraphics{ecm12.eps}}     
559:  & \scalebox{0.33}{\includegraphics{ecm16.eps}}\\
560:    \psfrag{MCPA}[B1][B1][2][0]{MCPA centre}\scalebox{0.33}{\includegraphics{mcpa8.eps}}    
561:  & \scalebox{0.33}{\includegraphics{mcpa12.eps}}   
562:  & \scalebox{0.33}{\includegraphics{mcpa16.eps}}\\ 
563:    \psfrag{MCPA}[B1][B1][2][0]{MCPA average}\scalebox{0.33}{\includegraphics{mcpa8avg.eps}} 
564:  & \scalebox{0.33}{\includegraphics{mcpa12avg.eps}} 
565:  & \scalebox{0.33}{\includegraphics{mcpa16avg.eps}}\\
566:    \psfrag{NLCPA}[B1][B1][2][0]{NLCPA periodic}\scalebox{0.33}{\includegraphics{tb8p.eps}}
567:  & \scalebox{0.33}{\includegraphics{tb12p.eps}}     
568:  & \scalebox{0.33}{\includegraphics{tb16p.eps}}\\
569:    \psfrag{NLCPA}[B1][B1][2][0]{NLCPA anti-periodic}\scalebox{0.33}{\includegraphics{tb8a.eps}} 
570:  & \scalebox{0.33}{\includegraphics{tb12a.eps}}     
571:  & \scalebox{0.33}{\includegraphics{tb16a.eps}}\\ 
572:    \psfrag{Nc8}[B1][B1][2.5][0]{$N_C=8$}\psfrag{NLCPA}[B1][B1][2][0]{NLCPA average}\scalebox{0.33}{\includegraphics{tb8avg.eps}}
573:  & \psfrag{Nc12}[B1][B1][2.5][0]{$N_C=12$}\scalebox{0.33}{\includegraphics{tb12avg.eps}}   
574:  & \psfrag{Nc16}[B1][B1][2.5][0]{$N_C=16$}\scalebox{0.33}{\includegraphics{tb16avg.eps}} 
575:  \end{tabular}         
576:  \caption{Configurationally-averaged DOS per site as a function of energy (in units of the bandwidth) for the various cluster theories (top 
577:  to bottom) with cluster sizes $N_c=8,12,16$ (left to right).}\label{cluster81216}
578:  \end{center}
579: \end{figure*}
580: 
581: The ability of the various cluster theories to improve upon the CPA result is now examined. All results here have been displayed using
582: centred histograms at the same energy resolution as the exact result shown in figure~\ref{pureexactcpa}(c), and all energies have a $10^{-3}$
583: imaginary part.
584: 
585: Figures~\ref{cluster246},\ref{cluster81216} show the DOS for the random~\footnote{An investigation of the ability of these cluster theories
586: to deal with the effects of chemical short-range order is deferred to another publication.} $A_{50}B_{50}$ alloy of figure~\ref{pureexactcpa}
587: using the various cluster theories (top to bottom) as a function of cluster size (left to right). For the ECM calculations shown in the first
588: row, the configurational average has been taken over the $2^{N_c}$ possible cluster configurations and the DOS measured on one of the central
589: sites. The central site is likely to give a better description of the alloy since it is more influenced by the effects of the disorder than
590: say a site on the boundary. Similarly, the MCPA calculations on the second row (`MCPA centre') have been measured at one of the central
591: sites. For the MCPA calculations on the third row (`MCPA average'), the DOS has been measured at each site on the cluster, and then an
592: average taken over all such measurements. The fourth and fifth rows show NLCPA calculations using the periodic and anti-periodic sets of
593: cluster momenta $\{\b{K}_n^{P}\}$ and $\{\b{K}_n^{A}\}$ respectively. Finally, the sixth row (`NLCPA average') shows an average of the
594: periodic and anti-periodic results given in the fourth and fifth rows.
595: 
596: First let us consider the ECM and MCPA central site results shown in the first two rows. As a general observation, it can be seen that all
597: plots show increasing structure in the DOS as the cluster size increases. This arises from statistical fluctuations in the local environment
598: of a site (smoothed out by the single-site mean-field CPA) due to specific impurity cluster configurations, and can be identified with
599: corresponding structure in the exact DOS shown in figure~\ref{pureexactcpa}(c). Furthermore, the gap at $E=0.0$ is systematically filled in
600: as $N_c$ increases due to larger cluster configurations present which contribute to the DOS at the band edges. Although this occurs more
601: quickly with the MCPA than the ECM, it is interesting to observe that the ECM results and the MCPA central site results are very similar for
602: all cluster sizes, indicating that self-consistency with respect to the cluster does not have a significantly large effect on the DOS results
603: for this system (in the absence of chemical SRO). Indeed, the ECM and MCPA central site DOSs become identical for $N_c=12$, and both converge
604: almost to the exact result at $N_c=16$.~\footnote{In theory, the MCPA has the incorrect boundary conditions as $N_c\rightarrow\infty$ since
605: the self-energy would become independent of $\b{k}$. In practice, however, the DOS converges quickly to the exact result.} Next, consider the
606: `MCPA average' results shown in the third row. For $N_c=2$, this is identical to the MCPA central site result by symmetry. As $N_c$
607: increases, more structure is again present in the DOS, although it is evident that the convergence towards the exact result is much slower.
608: For example, the troughs in the DOS deepen slightly as $N_c$ increases from 8 to 16, but a higher cluster size would be needed for the
609: troughs to reach the zero axis as seen in the exact result.
610: 
611: Now let us turn our attention to the NLCPA calculations. First consider $N_c=2$. In the periodic (P) result, the peaks in the DOS can be
612: identified with corresponding peaks present in the ECM and MCPA calculations, but the troughs present at $E=\pm{3}$ and $E=\pm{1.5}$ are not
613: seen elsewhere. The anti-periodic (AP) result is equivalent to the CPA here due to symmetry. On the other hand, the average over the P and AP
614: results shown in the sixth row resembles the MCPA result much more closely, with peaks and troughs lying at the same energies. For $N_c=4$,
615: both the P and AP calculations separately resemble the ECM and MCPA only over certain energy regions. However, the average over both the A
616: and AP results again closely resembles those of the ECM and MCPA. Thus already, it is clear that P and AP calculations separately are not
617: sufficient to yield a physical DOS, but the average does yield a physically meaningful DOS with peaks and troughs which can be associated
618: with correlations in the cluster disorder configurations. The reason behind this is clear from section~\ref{solutions}; the P and AP
619: calculations separately reproduce correlations arising from differing regions of reciprocal space, and thus a combination of the two
620: calculations is necessary to reproduce all correlations within the range of the cluster size. The more sophisticated method for combining the
621: A and AP solutions \cite{Batt1} mentioned in section~\ref{solutions} is likely to give even better results than the straight average
622: displayed here, for example the `flat' regions in the $N_c=2$ calculation are likely to be smoothed out. Also, observe that as $N_c$
623: increases further, the A and AP plots appear increasingly similar, becoming almost identical at $N_c=12$. This is expected since
624: $\{\b{K}_n^{P}\}$ and $\{\b{K}_n^{AP}\}$ become closer together as $N_c$ increases. 
625: 
626: Next, let us consider the convergence properties of the NLCPA. Importantly, it is clear that, provided the periodic and anti-periodic results
627: are combined, the DOS calculated using the NLCPA converges smoothly and systematically towards the exact result. There are however clear
628: differences in the convergence properties of the NLCPA compared to the ECM and MCPA. Firstly, the gap centred at $E=0.0$ is filled in more
629: quickly. Secondly, the convergence for the NLCPA is more systematic in that the maximum height of the DOS increases monotonically as $N_c$
630: increases ($N_c=20$ and $N_c=24$ calculations further confirm this). On the other hand, this means that large `spikes' present in the ECM and
631: MCPA DOS are not reproduced for small cluster sizes, although such spikes are seen in the exact result. Indeed, the ECM and MCPA plots show
632: much more fine structure for a given cluster size, whilst the NLCPA results are much more `smeared out'. Consequently, the ECM and MCPA
633: central site calculations converge much more quickly to the exact result as $N_c$ increases. However, it is important to appreciate that the
634: main (and more important) features of the exact result are reproduced fairly quickly by the NLCPA. Also, the ECM and MCPA calculations are
635: very sensitive to the energy resolution at which they are displayed, unlike the NLCPA. These differences in the convergence properties are
636: all a consequence of the fact that the NLCPA is based on Nyquist's sampling theorem \cite{Elliot1,Jarrell1}. Finally, observe that both the
637: MCPA average (third row in figures~\ref{cluster246},\ref{cluster81216}) and the NLCPA have not yet converged to the exact result at $N_c=16$.
638: Indeed there has been much debate in the literature~\cite{Maier1,Biroli1,Aryanpour1,Biroli2} concerning whether the dynamical cluster
639: approximation (DCA)~\cite{Hettler1} or the cellular dynamical mean field theory (CDMFT)~\cite{Kotliar1} (the analogues of the NLCPA and MCPA
640: respectively for strongly-correlated electron systems) converges more quickly, carried out in terms of an average hybridization function
641: coupling the cluster to the medium. Whilst a convergence comparison of the DOS using the NLCPA and MCPA average is not one of the objectives
642: of this paper, it would be interesting to carry this out by using larger (and computationally very demanding) cluster sizes. It is clear,
643: however, that a convergence comparison depends greatly on whether the quantity in question is sensitive to the energy resolution. For
644: example, the convergence properties are likely to be much more similar for a quantity such as the integrated DOS.
645: 
646: %--------------------------------------------------------------------------------------------------------------------------------------------
647: \section{Conclusions}\label{conclude}
648: 
649: A comparison of the formalism of the NLCPA with the ECM and MCPA has been presented. In the ECM and MCPA, the cavity function introduced has
650: an explicit real-space expansion and so the coupling of an impurity cluster to the surrounding medium is well defined. It is found that this
651: is not the case for the NLCPA cavity function (for $N_c>1$) as a consequence of the introduction of Born-von Karman boundary conditions. This
652: implies that the cavity should be viewed as a mathematical construction used to determine the medium, in departure from traditional effective
653: medium theories. The consequences of introducing anti-periodic Born-von Karman boundary conditions within the NLCPA has also been
654: investigated. Numerical DOS results for a 1D model have been compared. It can be concluded that, firstly, the periodic and anti-periodic
655: solutions must be combined when using the NLCPA, particularly for small cluster sizes. Importantly, only by then comparing the DOS with the
656: exact result does it become evident that the NLCPA does indeed produce physical results and, for each cluster size, shares main features in
657: common with both the ECM and MCPA. Since the NLCPA is based upon Nyquist's sampling theorem, it is however found that the DOS is far more
658: `smeared out' for a given cluster size. Whilst the NLCPA DOS does (importantly) converge systematically towards the exact result as the
659: cluster size increases, it therefore does so less quickly than the ECM and MCPA (measured on the central site and in the absence of chemical 
660: SRO). On the other hand, such fine structure is less important and in any case would be smeared out in higher dimensions. More importantly, 
661: the NLCPA preserves the full symmetry of the underlying lattice, yields well-defined site-diagonal and site off-diagonal properties, and is 
662: computationally tractable for realistic systems.
663: 
664: %--------------------------------------------------------------------------------------------------------------------------------------------
665: \ack
666: 
667: The author thanks G.~M.~Batt, A.~Gonis, B.~L.~Gy{\"{o}}rffy, and M.~Jarrell for useful discussions, and D.~K{\"{o}}dderitszch for the exact
668: histogram results. This work was funded by a research fellowship in theoretical physics from EPSRC~(UK), grant GR/S92212/01.
669: 
670: \section*{References}
671: 
672: \begin{thebibliography}{10}
673: 
674: \bibitem{Soven1}
675: P. Soven, Phys. Rev. {\bf 156},  809  (1967).
676: 
677: \bibitem{Gonis1}
678: A. Gonis, {\em Green Functions for Ordered and Disordered Systems}, Vol.~4 of
679:   {\em Studies in Mathematical Physics} (North Holland, Amsterdam, 1992).
680: 
681: \bibitem{Jarrell1}
682: M. Jarrell and H.~R. Krishnamurthy, Phys. Rev. B {\bf 63},  125102  (2001).
683: 
684: \bibitem{Gonis6}
685: A. Gonis and J.~W. Garland, Phys. Rev. B {\bf 16},  2424  (1977).
686: 
687: \bibitem{Tsukada1}
688: M. Tsukada, J. Phys. Soc. Jpn {\bf 32},  1475  (1972).
689: 
690: \bibitem{Hettler1}
691: M.~H. Hettler {\it et~al.}, Phys. Rev B. {\bf 58},  7475  (1998).
692: 
693: \bibitem{Hettler2}
694: M.~H. Hettler, M. Mukherjee, M. Jarrell, and H.~R. Krishnamurthy, Phys. Rev. B
695:   {\bf 61},  12739  (2000).
696: 
697: \bibitem{Moradian2}
698: R. Moradian, B.~L. Gy{\"{o}}rffy, and J.~F. Annett, Phys. Rev. Lett. {\bf 89},
699:   287002  (2002).
700: 
701: \bibitem{Rowlands1}
702: D.~A.~Rowlands, {\it{`The KKR-NLCPA: a new method for describing the electronic
703:   structure of disordered metallic sytems'}}, Ph.D.~Thesis, University of
704:   Warwick (2004).
705: 
706: \bibitem{Korringa1}
707: J. Korringa, Physica {\bf 13},  392  (1947).
708: 
709: \bibitem{Kohn2}
710: W. Kohn and N. Rostoker, Phys. Rev. {\bf 94},  1111  (1954).
711: 
712: \bibitem{Rowlands2}
713: D.~A. Rowlands, J.~B. Staunton, and B.~L. Gy{\"{o}}rffy, Phys. Rev. B. {\bf
714:   67},  115109  (2003).
715: 
716: \bibitem{Rowlands3}
717: D.~A.~Rowlands, J.~B.~Staunton, B.~L.~Gy{\"{o}}rffy, E.~Bruno, and
718:   B.~Ginatempo, cond-mat/0411347, Phys.~Rev.~B \textbf{72}, 045101 (2005).
719: 
720: \bibitem{Rowlands5}
721: D.~A.~Rowlands, A.~Ernst, B.~L.~Gy{\"{o}}rffy, and J.~B.~Staunton; to be
722:   published.
723: 
724: \bibitem{Staunton2}
725: J.~B. Staunton and B.~L. Gy{\"{o}}rffy, Phys. Rev. Lett. {\bf 69},  371
726:   (1992).
727: 
728: \bibitem{Gyorffy9}
729: B.~L. Gy{\"{o}}rffy, Physica Scripta {\bf T49},  373  (1993).
730: 
731: \bibitem{Luders1}
732: M. L{\"{u}}ders {\it et~al.}, Phys. Rev. B {\bf 71},  205109  (2005).
733: 
734: \bibitem{Ashcroft1}
735: N.~W. Ashcroft and N.~D. Mermin, {\em Solid State Physics} (Holt, Rinehart and
736:   Winston, New York, 1976).
737: 
738: \bibitem{Elliot1}
739: D.~F. Elliot and K.~R. Rao, {\em Fast Transforms: Algorithms, Analyses,
740:   Applications} (Academic, New York, 1982).
741: 
742: \bibitem{Gonis5}
743: A. Gonis,  in {\em Theoretical Materials Science: Tracing the Origins of
744:   Materials Behaviour} (Materials Research Society, Warrendale, 2000), p.\ 365.
745: 
746: \bibitem{Maier3}
747: T.~A. Maier, T. Pruschke, and M. Jarrell, Phys. Rev. B {\bf 66},  075102
748:   (2002).
749: 
750: \bibitem{Batt1}
751: G.~M.~Batt, B.~L.~Gy{\"{o}}rffy, and D.~A.~Rowlands; to be published.
752: 
753: \bibitem{Jarrell2}
754: M.~Jarrell; private communication (2003).
755: 
756: \bibitem{Moradian1}
757: R.~Moradian, Ph.D.~Thesis, University of Bristol (2001).
758: 
759: \bibitem{Ziman1}
760: J.~M. Ziman, {\em Models of Disorder} (Cambridge University Press, Cambridge,
761:   1979).
762: 
763: \bibitem{Dean1}
764: P. Dean, Rev. Mod. Phys. {\bf 44},  127  (1972).
765: 
766: \bibitem{Maier1}
767: T.~A. Maier and M. Jarrell, Phys. Rev. B {\bf 65},  041104  (2002).
768: 
769: \bibitem{Biroli1}
770: G. Biroli and G. Kotliar, Phys. Rev. B {\bf 65},  155112  (2001).
771: 
772: \bibitem{Aryanpour1}
773: K. Aryanpour, T.~A. Maier, and M. Jarrell, Phys. Rev. B {\bf 71},  037101
774:   (2005).
775: 
776: \bibitem{Biroli2}
777: G. Biroli and G. Kotliar, Phys. Rev. B {\bf 71},  037102  (2005).
778: 
779: \bibitem{Kotliar1}
780: G. Kotliar, S.~Y. Savrasov, G. Pálsson, and G. Biroli, Phys. Rev. Lett. {\bf
781:   87},  186401  (2001).
782: 
783: \end{thebibliography}
784: 
785: \end{document}
786: 
787: 
788: 
789: 
790: