1: %%%%%%%%%%%%%%%%%%%%%%% file template.tex %%%%%%%%%%%%%%%%%%%%%%%%%
2: %
3: % This is a template file for The European Physical Journal
4: %
5: % Copy it to a new file with a new name and use it as the basis
6: % for your article
7: %
8: %%%%%%%%%%%%%%%%%%%%%%%% Springer-Verlag %%%%%%%%%%%%%%%%%%%%%%%%%%
9: %
10: \begin{filecontents}{leer.eps}
11: %!PS-Adobe-2.0 EPSF-2.0
12: %%CreationDate: Mon Jul 13 16:51:17 1992
13: %%DocumentFonts: (atend)
14: %%Pages: 0 1
15: %%BoundingBox: 72 31 601 342
16: %%EndComments
17:
18: gsave
19: 72 31 moveto
20: 72 342 lineto
21: 601 342 lineto
22: 601 31 lineto
23: 72 31 lineto
24: showpage
25: grestore
26: %%Trailer
27: %%DocumentFonts: Helvetica
28: \end{filecontents}
29: %
30: \documentclass[epj]{svjour}
31: % Remove option referee for final version
32: %
33: % Remove any % below to load the required packages
34: %\usepackage{latexsym}
35: \usepackage{graphics}
36: % etc
37: %
38: \begin{document}
39: %
40: \title{CPA density of states and conductivity in a
41: double-exchange system containing impurities}
42: \titlerunning{CPA in a double-exchange system}
43: \author{Mark Auslender\inst{1}\thanks{\emph{e-mail:}
44: marka@ee.bgu.ac.il}
45: \and Eugene Kogan\inst{2} \thanks{\emph{e-mail:}
46: kogan@quantum.ph.biu.ac.il}} % Do not remove
47: %
48: %\offprints{} % Insert a name or remove this line
49: %
50: \institute{Department of Electrical and Computer Engineering,
51: Ben-Gurion University of the Negev P.O.B. 653, Beer-Sheva 84105, Israel
52: \and Jack and Pearl Resnick Institute of Advanced Technology,
53: Department of Physics, Bar-Ilan University, Ramat-Gan 52900, Israel }
54: \authorrunning{Auslender and Kogan }
55: %
56: \date{Received: date / Revised version: date}
57: % The correct dates will be entered by Springer
58: %
59: \abstract
60: {We study density of states and conductivity of the doped double-exchange
61: system, treating interaction of charge carriers both with the
62: localized spins
63: and with the impurities in
64: the coherent potential approximation. It is shown that under appropriate
65: conditions
66: there is a gap between the
67: conduction band and the impurity band in paramagnetic phase, while
68: the density of states is gapless in ferromagnetic phase.
69: This can explain metal-insulator transition frequently observed in manganites and
70: magnetic semiconductors. Activated conductivity in the insulator phase is
71: numerically calculated.
72: \PACS{
73: {75.50.Pp} {Magnetic semiconductors} \and
74: {75.30.Vn} {Colossal magnetoresistance} \and
75: {72.10.-d} {Theory of electronic transport; scattering mechanisms}
76: } % end of PACS codes
77: } %end of abstract
78: %
79: \maketitle
80: %
81:
82:
83: \section{Introduction}
84:
85: The recent rediscovery of colossal magnetoresistance (CMR) in doped Mn oxides
86: with perovskite structure R$_{1-x}$D$_x$MnO$_3$ (R is a rare-earth metal and D
87: is a divalent metal, typically Ba, Sr or Ca) \cite{helmolt} has
88: generated substantial interest in these materials \cite{ramirez}.
89: The doping of parent material RMnO$_3$ by a divalent metal is the source of
90: the holes responsible for the transport properties of these materials.
91: In addition, each divalent atom introduced, is the center of an impurity
92: potential. Many papers analyzed the influence of strong magnetic disorder, inherent in the
93: the CMR materials at finite temperature, upon the single-particle states and transport
94: properties. However, the interplay between the magnetic disorder and the doping-induced
95: disorder was studied less. The impurity potential plays double role.
96: First, the potential fluctuations determine the transport at temperatures well
97: below the ferromagnet (FM) - paramagnet (PM) transition point $T_c$. Second,
98: strong potential may pin the
99: Fermi level either in the conduction band tail
100: (in the Anderson model
101: of disorder \cite{sheng}), or in the emerging impurity band.
102: The analysis of experimental data
103: reveals strong relevance of the latter effect to metal-insulator transition (MIT) near $T_c$
104: both in magnetic semiconductors \cite{kog} and manganites \cite{bebenin}.
105: However, to the best
106: of our knowledge the impurity-band scenario in the double-exchange (DE)
107: model was not discussed yet.
108:
109: The present paper is devoted to the consideration of single-particle states and
110: conductivity in impure DE system. Interaction of charge carriers both with the
111: localized spins and with the impurities is strong, so it is definitely not enough to
112: limit ourselves with the finite number of terms of
113: perturbation expansion. A simple but physically meaningful approximation,
114: allowing to sum up
115: infinite number of
116: perturbation expansion terms is the
117: coherent potential approximation (CPA). Initially CPA was proposed
118: to treat potential disorder \cite{ziman}, but soon after it's appearance the
119: generalization to random spin system was developed \cite{kubo}. The CPA was also
120: used to describe diluted magnetic semiconductors \cite{takahashi}.
121:
122: In the present paper we for the first time treat on equal footing the
123: interactions of electrons with the core Mn spins and with the
124: doping impurities using
125: the matrix generalization of the CPA. The concurrent action of
126: potential disorder and temperature dependent spin disorder leads to a number of
127: interesting phenomena, in particular to the possibility of the opening of the
128: gap at the Fermi level with the increase of temperature and, hence, to
129: MIT transition.
130:
131:
132: \section{Hamiltonian and Theoretical Formulation}
133:
134: We consider the DE model with the inclusion of the
135: single-site impurity potential. In addition, as it is widely accepted,
136: we apply the quasiclassical adiabatic approximation and consider each
137: Mn spin as a static vector
138: with a fixed length $S$ (${\bf S}_i=S{\bf n}_i$, where ${\bf n}_i$
139: is a randomly oriented unit vector).
140: The Hamiltonian of the model in site representation is
141: \begin{eqnarray}
142: \label{ham}
143: \hat{H}_{ij}=t_{i-j}+
144: \delta_{ij}\left(\epsilon_i
145: -J{\bf n}_i\cdot\hat{\bf \sigma}\right)=H_{kin}+V_{imp}+\hat{V}_{sd},
146: \end{eqnarray}
147: where $t_{i-j}$ is the electron hopping, $\epsilon_i$ is the
148: random on-site energy, $J$ is the effective
149: exchange
150: coupling between a Mn core spin and a conduction electron and $
151: \hat{\bf \sigma}$ is the vector of the Pauli matrices. The hat above the
152: operator reminds that in one-particle representation it is a
153: $2\times 2$ matrix in the spin space (we discard the
154: hat when the operator is a scalar matrix in the spin space).
155:
156: We present Hamiltonian as
157: \begin{equation}
158: \hat{H}=H_{kin}+\hat{\Sigma}+V_{imp}+\hat{V}_{sd}-\hat{\Sigma}=
159: \hat{H}_0+\hat{V}
160: \end{equation}
161: (the site independent self-energy $\hat{\Sigma}(E)$ is to be determined later),
162: and construct a perturbation theory with respect to random potential
163: $\hat{V}=V_{imp}+\hat{V}_{sd}-\hat{\Sigma}$.
164: To do this let us
165: introduce the $T$-matrix as the solution of the equation
166: \begin{equation}
167: \hat{T}=\hat{V}+\hat{V}\hat{G}_0\hat{T},
168: \end{equation}
169: where
170: \begin{equation}
171: \label{efes}
172: \hat{G}_0=\frac{1}{E-\hat{H}_0}.
173: \end{equation}
174: For the exact Green function we get
175: \begin{equation}
176: \label{green}
177: \hat{G}=\hat{G}_0+\hat{G}_0\hat{T}\hat{G}_0.
178: \end{equation}
179: The coherent potential approximation (CPA)
180: is expressed by the equation
181: \begin{equation}
182: \label{g}
183: \left\langle\hat{G}\right\rangle=\hat{G}_0.
184: \end{equation}
185: This equation can also be presented as
186: \begin{equation}
187: \label{t}
188: \left\langle\hat{T}_i\right\rangle=0,
189: \end{equation}
190: where $\hat{T}_i$ is the solution of the equation
191: \begin{equation}
192: \hat{T}_i=\hat{V}_i+\hat{V}_i
193: \hat{g}(E-\hat{\Sigma})\hat{T}_i,
194: \end{equation}
195: and
196: \begin{equation}
197: \label{locator}
198: g(E)=\left(G_0(E)\right)_{ii}=
199: \int\frac{N_0(\varepsilon)}{E-\varepsilon}d\varepsilon,
200: \end{equation}
201: where $N_0(\varepsilon)$ is the bare density of states.
202: The averaging in Eqs. (\ref{g},\ref{t}) should be performed both with respect to
203: random orientations of core spins and with respect to random on-site energies.
204: We obtained, in fact, the algebraic equation for the $2\times 2$ matrix $\hat{\Sigma}$
205: \begin{equation}
206: \label{gencpa}
207: \left\langle\left[1-\hat{V}_i\hat{g}(E-\hat{\Sigma})\right]^{-1}
208: \hat{V}_i\right\rangle=0.
209: \end{equation}
210: This equation takes into account scattering both due to randomness of the core
211: spins, and due to the impurities.
212: If the impurity potential is negligible
213: ($V=0$) this equation coincides with the Eq.(20) of Ref.
214: \cite{furukawa} obtained in the dynamical mean field approximation (and also
215: with those obtained for the
216: Falikov-Kimball model \cite{moller,DMFA}).
217:
218: In the reference frame where the $z$ axis is directed along the magnetization,
219: $\hat{\Sigma}$
220: is diagonal, and Eq.(\ref{gencpa}) reduces to the system of two equations for
221: its diagonal
222: matrix elements $\Sigma_{\sigma}(E)$ ($\sigma = \uparrow,\downarrow$).
223: The equations acquire especially simple form at two extreme
224: particular cases,which we will analyze:
225:
226: \noindent (a) $T=0$. The magnetic state is coherent FM with $n_i^z=1$, and
227: Eq. (\ref{gencpa}) takes the form
228: \begin{equation}
229: \left\langle \frac{\epsilon_i \mp J-\Sigma _{\uparrow ,\downarrow }}
230: {1-\left( \epsilon_i \mp J-\Sigma _{\uparrow ,\downarrow }\right)
231: g(E-\Sigma_{\uparrow ,\downarrow })}\right\rangle =0 \label{ztcpa}.
232: \end{equation}
233:
234: \noindent (b) $T\geq T_{c}$ and zero magnetic field.
235: The magnetic state is isotropic PM with
236: $\langle {\bf n_{i}}\rangle =0$, which leads to
237: $\Sigma _{\downarrow }=\Sigma_{\uparrow }=\Sigma $, and
238: Eq. (\ref{gencpa}) takes the form.
239: \begin{eqnarray}
240: \label{htcpa}
241: \left\langle\frac{\epsilon_i +J-\Sigma }{1-\left( \epsilon_i +J-\Sigma\right)
242: g(E-\Sigma)}\right\rangle \nonumber\\+
243: \left\langle\frac{\epsilon_i -J-\Sigma }
244: {1-\left(\epsilon_i -J-\Sigma \right) g(E-\Sigma)}\right \rangle = 0.
245: \end{eqnarray}
246:
247: We will solve the equations Eqs.(\ref{ztcpa}) and (\ref{htcpa}) in
248: the strong Hund coupling limit
249: ($J\rightarrow \infty $).
250: In this limit we obtain two decoupled spin sub-bands. The equation for the upper
251: sub-band, after shifting the energy by $-J$, for both cases (a) and (b) can be written
252: down in unified form
253: \begin{equation}
254: \left\langle \frac{1}{1-\left( \epsilon_i -\Sigma \right)
255: g( E -\Sigma) }\right\rangle =\alpha,
256: \label{infJcpa}
257: \end{equation}
258: where $\alpha =1$ for $T=0$, $\alpha =2$ for $T\geq T_{c}$.
259: In the model of substitutional
260: disorder ($\epsilon_i = 0$ with probability $x$, and $\epsilon_i =V$ with probability
261: $1-x$),
262: Eq.(\ref{infJcpa}) takes the form
263: \begin{equation}
264: \frac{1-x}{1+\Sigma g\left(E - \Sigma \right) }+
265: \frac{x}{1+\left(\Sigma -V\right) g\left( E -\Sigma\right) }=\alpha.
266: \label{infJcpasd}
267: \end{equation}
268:
269: \section{The CPA equations for semi-circular bare density of states}
270:
271:
272: We consider semi-circular (SC) bare DOS
273: \begin{equation}
274: N_{0}( \varepsilon) =\frac{4}{\pi W}
275: \sqrt{1-\left( \frac{2\varepsilon}{W}\right) ^{2}},
276: \label{scdos}
277: \end{equation}
278: at $\left| \varepsilon\right| \leq W/2$ and $N_{0}(\varepsilon) =0$
279: otherwise, for
280: which
281: \begin{equation}
282: g(E) =\frac{4}{W}\left[\frac{2E}{W}-\sqrt{\left( \frac{2E}{W}\right)^{2}-1}\right].
283: \label{scgrfun}
284: \end{equation}
285: Let us introduce the following normalized quantities
286: \begin{eqnarray}
287: \lambda=\frac{\Sigma}{W},\;\omega =\frac{E}{W},\;v=\frac{V}{W}
288: \label{dimlessgf}
289: \end{eqnarray}
290: After simple algebra we obtain from Eq. (\ref{infJcpasd})
291: the cubic equation with respect to
292: \begin{eqnarray}
293: \gamma \equiv Wg\left(E -\Sigma \right) =8\left[ \omega -\lambda -
294: \sqrt{\left( \omega -\lambda \right) ^{2}-1/4}\right],
295: \end{eqnarray}
296: in the form
297: \begin{eqnarray}
298: \gamma ^{3}+16\left( v-2\omega \right) \gamma ^{2}+16
299: \left[\frac{1}{\alpha} -16\omega \left( v-\omega \right) \right]
300: \gamma\nonumber\\
301: -256\frac{\omega}{\alpha}
302: +256\left( 1-x\right) \frac{v}{\alpha} =0 \label{cubicgam}.
303: \end{eqnarray}
304: The number of electrons per cite $n$ is given by
305: \begin{equation}
306: \label{fermi}
307: n=\int_{-\infty}^{\infty}f(E)N(E)dE,
308: \end{equation}
309: where $f(E)$ is the Fermi distribution function, and
310: \begin{equation}
311: N(E)=\frac{\alpha}{W\pi}\mbox{Im}\;\gamma
312: \end{equation}
313: is the actual density of states.
314: To define the position of $\mu$, the Fermi level,
315: we must impose the relation
316: between $n$ and $x$; the simplest assumption appropriate for manganites is the
317: equation $n=1-x$.
318:
319: \section{Conductivity in CPA}
320:
321: For a disordered one-electron system the static conductivity is given by
322:
323: \begin{eqnarray}
324: \rho^{-1} =\frac{e^{2}\pi \hbar }{\rm V}
325: \int \left(-\frac{\partial f}{\partial E}\right)\nonumber\\
326: \cdot\left\langle \mbox{Tr}\left[ \hat{v}_{\alpha }\delta \left(E-\hat{H}\right)
327: \hat{v}_{\alpha }\delta \left( E-\hat{H}\right) \right]
328: \right\rangle dE \label{kubogrin},
329: \end{eqnarray}
330: where ${\rm V} $ is the volume and $\hat{v}_{a}$ is a Cartesian component of
331: the velocity operator. To
332: obtain the conductivity in CPA let us express operator delta-function as
333: follows
334: \begin{equation}
335: \delta \left( E-\hat{H}\right) =\frac{1}{2\pi i}\left[ \hat{G}%
336: (E_{-})-\hat{G}(E_{+})\right].
337: \label{specop}
338: \end{equation}
339: Using Eq. (\ref{green}) and Eq. (\ref{efes}) in Bloch representation
340: \begin{equation}
341: \left\langle {\bf k}\sigma \left| \hat{G}_{0}(E)\right| {\bf k}%
342: ^{\prime }\sigma ^{\prime }\right\rangle =\frac{\delta _{\mathbf{k,k}%
343: ^{\prime }}\delta _{\sigma,\sigma^{\prime }}}{E
344: -\varepsilon_{\mathbf{k}}-\Sigma _{\sigma }\left( E\right) },
345: \label{grinfun}
346: \end{equation}
347: we get
348: \begin{eqnarray}
349: \left\langle \mbox{Tr}\left[ \hat{v}_{a}\delta \left( E-\hat{H}%
350: \right) \hat{v}_{a}\delta \left( E-\hat{H}\right) \right]
351: \right\rangle \nonumber\\
352: =\sum_{\mathbf{k},\sigma }v_{\mathbf{k}\alpha }^{2}\left[
353: A_{\sigma }\left( \varepsilon_{\mathbf{k}},E\right) \right] ^{2}+
354: O(\left\langle \hat{T}\hat{T}\right\rangle),
355: \end{eqnarray}
356: where
357: \begin{equation}
358: A_{\sigma }\left( \varepsilon ,E\right)
359: =\frac{1}{\pi}\frac{\mbox{Im}\Sigma _{\sigma
360: }\left( E\right) }{\left[ E-\varepsilon -\mbox{Re}\Sigma _{\sigma }\left(
361: E\right) \right] ^{2}+\left[ \mbox{Im}\Sigma _{\sigma }\left( E\right)
362: \right] ^{2}} \label{speccpa}
363: \end{equation}
364: is the one-particle spectral weight function. On account of the locality
365: of $T$-matrix the second term in the trace is equal to
366: \begin{eqnarray}
367: \sum_{s,s^{^{\prime }}=\pm;\;
368: \mathbf{k,k}^{\prime },\sigma ,\sigma^{\prime }}
369: ss^{\prime }v_{\mathbf{k}\alpha }v_{\mathbf{k}^{\prime }\alpha
370: }G_{\sigma }\left( \varepsilon_{\mathbf{k}},E_{s}\right) G_{\sigma ^{\prime
371: }}\left( \varepsilon_{\mathbf{k}^{\prime }},E_{s}\right)\nonumber\\
372: \times G_{\sigma ^{\prime }}\left( \varepsilon_{\mathbf{k}^{\prime
373: }},E_{s^{\prime }}\right) G_{\sigma }\left( \varepsilon_{\mathbf{k}%
374: },E_{s^{\prime }}\right) \nonumber\\
375: \times \left\langle T_{\sigma \sigma ^{\prime }}
376: ({\bf k}-{\bf k}^{\prime },E_{s})T_{\sigma ^{\prime }\sigma }
377: ({\bf k}^{\prime }-{\bf k},E_{s^{\prime }})\right\rangle.
378: \end{eqnarray}
379: Since in CPA $\left\langle T_{\sigma \sigma ^{\prime }}
380: ({\bf k}-{\bf k}^{\prime },E_{s})T_{\sigma ^{\prime }\sigma }
381: ({\bf k}^{\prime }-{\bf k},E_{s^{\prime }})\right\rangle$ does not depend on
382: $\mathbf{k}$ and $\mathbf{k}^{\prime}$
383: and $v_{-\mathbf{k}\alpha}=-v_{\mathbf{k}\alpha}$ the above expression is identically
384: zero \cite{velic,khurana}.
385: Thus, finally
386: \begin{eqnarray}
387: \rho^{-1} =\frac{e^{2}\pi \hbar }{ v }\int
388: \int \left(-\frac{\partial f}{\partial E}\right)\nonumber\\
389: \cdot v_{\alpha }^{2}\left( \varepsilon \right) N_{0}\left( \varepsilon \right)
390: \sum_{\sigma }\left[ A_{\sigma }\left( \varepsilon ,E\right) \right]
391: ^{2}dEd\varepsilon,
392: \label{cpacond}
393: \end{eqnarray}
394: where $v$ is the unit cell volume and by definition
395: \begin{equation}
396: v_{\alpha}^{2}\left( \varepsilon \right) N_{0}\left( \varepsilon \right)
397: =\frac{1}{N\hbar ^{2}}
398: \sum_{\mathbf{k}}\left( \frac{\partial \varepsilon_{\mathbf{k}}}
399: {\partial k_{\alpha }}\right) ^{2}\delta \left( \varepsilon -\varepsilon
400: _{\mathbf{k}}\right).
401: \label{msqvel}
402: \end{equation}
403:
404: Let us assume nearest-neighbor tight binding spectrum on simple $d$%
405: -hypercubic lattice ($v =a^{d}$)
406: \begin{eqnarray}
407: \varepsilon_{\mathbf{k}} = -t\sum_{\alpha =1}^{d}\cos ak_{\alpha }, \nonumber \\
408: v_{\alpha}^{2}\left(\varepsilon \right) N_{0}\left( \varepsilon \right)
409: =-\frac{a^{2}}{d\hbar
410: ^{2}}\int_{-\infty }^{\varepsilon}\epsilon N_{0}\left( \epsilon \right)
411: d\epsilon.
412: \label{nnres}
413: \end{eqnarray}
414: In SC DOS model
415: \begin{eqnarray}
416: v_{\alpha }^{2}\left( \varepsilon \right) N_{0}\left( \varepsilon \right) =-
417: \frac{4}{\pi W}\frac{a^{2}}{d\hbar ^{2}}\int_{-W/2}^{\varepsilon }z\sqrt{
418: 1-\left( \frac{2z}{W}\right) ^{2}}dz\nonumber\\
419: =\frac{1}{3}\frac{W}{\pi }\frac{a^{2}}{%
420: d\hbar ^{2}}\left( 1-\frac{4\varepsilon ^{2}}{W^{2}}\right) ^{3/2}.
421: \end{eqnarray}
422: Substituting this result into Eq.(\ref{cpacond}) we obtain
423: \begin{equation}
424: \label{resist}
425: \rho ^{-1}=\sigma _{0}\int \left(-\frac{\partial f}{\partial E}\right)
426: \Lambda \left(E\right) dE,
427: \end{equation}
428: with
429: \begin{equation}
430: \sigma _{0}=\frac{e^{2}}{2\pi da^{d-2}\hbar }
431: \end{equation}
432: being the Mott minimal metallic conductivity, and
433: \begin{eqnarray}
434: \Lambda \left( E\right) =\frac{2W}{3\pi}\int_{-W/2}^{W/2} \left(
435: 1-\frac{4\varepsilon ^{2}}{W^{2}}\right) ^{3/2}\nonumber\\
436: \sum_{\sigma}\left\{ \mbox{Im}\left[ \frac{1}
437: {E-\varepsilon -\Sigma _{\sigma }\left( E\right) }\right] \right\}
438: ^{2}d\varepsilon.
439: \end{eqnarray}
440: For the strong Hund coupling we obtain
441: \begin{eqnarray}
442: \label{econd}
443: \Lambda \left( E\right) =
444: =\frac{4}{3\pi}\int_{-1}^{1}\left( 1-x^{2}\right) ^{3/2}
445: \left\{ \mbox{Im}\left[ \frac{1}{x-z}\right] \right\} ^{2}dx,
446: \end{eqnarray}
447: where $z = 2\left( \omega - \lambda \right)$.
448: Eqs. (\ref{resist}), (\ref{econd})
449: give the conductivity in the framework of bare SC DOS model
450: for arbitrary hole concentration $x$
451: and impurity potential strength $V$.
452:
453: \section{Influence of the impurity potential}
454:
455: First, consider density of states.
456: It is known, that within CPA for every $x$ there exists critical value of potential-to-bandwidth
457: ratio $v_c(x)$ such that at $v > v_c(x)$ the separate impurity band splits off the conduction band
458: (that is a gap opens in $N(E)$). In
459: our approach
460: we get two different curves $v_c(x,T=0)$ and $v_c(x,T \geq T_c)$, which present boundaries of
461: metal-insulator and metal-semiconductor 'phase diagrams', respectively,
462: in $(v,x)$ plane. Due to
463: effect of magnetic disorder it appears that $v_c(x,T=0) > v_c(x,T \geq T_c)$.
464:
465: For a typical concentration $x=0.2$
466: $v_{c}(0.2, T=0) \approx 0.49$ and $v_{c}(0.2, T\geq T_c) \approx 0.35$.
467: So if we choose
468: $v = 0.4$ both $N(E)$ and $\Lambda (E)$ must be gapless at $T=0$ but do have a gap at $T\geq T_c$.
469: Numerical calculations of the DOS performed at $T=0$ and $T\geq T_c$ for the above
470: $x$ and $v$ clearly demonstrate FM-PM transition induced band
471: splitting (Fig.1).
472:
473: \begin{figure}
474: \resizebox{0.4\textwidth}{!}{%
475: \includegraphics{fig1.eps}
476: }
477: \caption{DOS at $T=0$ (full line) and high-temperature (dashed line)
478: in the units of $W^{-1}$ for $x=0.2$ and $V/W=0.4$}
479: \label{fig:1} % Give a unique label
480: \end{figure}
481:
482: Now address the question of conductivity.
483: Consider first the position of $\mu$ and conductivity at $T=0$.
484: We get from Eq.(\ref{fermi})
485: $\mu(T=0)=0.3662W$. Note that $\mu(T=0)$ lies on the neck connecting
486: conduction band and impurity states derived parts of the band. As a result,
487: the residual conductivity (Eq.(\ref{resist}))
488: $\rho^{-1}(T=0)=0.6163\sigma_0$ is less then the Mott limit.
489:
490:
491: \begin{figure}
492: \resizebox{0.4\textwidth}{!}{%
493: \includegraphics{fig2.eps}
494: }
495: \caption{Energy-dependent conductivity at $T=0$ (full line)
496: and high-temperature
497: (dashed line) in the units of $\sigma_0$ for $x=0.2$ and $V/W=0.4$}
498: \label{fig:2} % Give a unique label
499: \end{figure}
500:
501:
502: At $T\geq T_c$ DOS and $\Lambda(E)$
503: have the same gap $\Delta=0.031W$ (see Figs.1,2), so $\mu(T\geq T_c)$ must lie
504: in the gap.
505: Thus, the model describes a bad metal at
506: $T=0$, and a semiconductor
507: at $T \geq T_c$. The transition between two types of
508: conduction (FM-PM
509: transition induced MIT) should occur at some temperature below $T_c$. Such a
510: picture agrees with the
511: recent photoemission experiments showing drastic decrease of DOS at the Fermi
512: level \cite{photo}
513: as temperature increases towards $T_c$.
514:
515: It is checked numerically that DOS displays square-root like behavior near the top of the
516: conduction band $E_{c}$
517: \begin{eqnarray}
518: N(E) \approx n_c\sqrt{E_{c}-E},
519: \end{eqnarray}
520: and the bottom of the impurity band $E_{i}$
521: \begin{eqnarray}
522: N(E) \approx n_i\sqrt{E-E_{i}}.
523: \end{eqnarray}
524: Unlike DOS $\Lambda (E)$ behaves {\it linearly} near the band edges
525: \begin{eqnarray}
526: \Lambda(E) \approx W^{-1}\lambda_c(E_{c}-E),\;\;\;\mbox{for}\; E<E_{c};\nonumber\\
527: \Lambda(E) \approx W^{-1}\lambda_i(E-E_{i}), \;\;\;\mbox{for}\; E>E_{i}.
528: \end{eqnarray}
529: The assumption $T<\Delta$
530: allows us to explicitly obtain $\mu(T\geq T_c)$.
531: Calculating integrals in
532: Eq.(\ref{fermi}) with exponential accuracy we obtain
533: \begin{equation}
534: \mu \approx \frac{1}{2}\left( E_{c}+E_{i}+T\ln\frac{n_c}{n_i}\right).
535: \end{equation}
536: The integral in Eq.(\ref{resist}), calculated with the same accuracy,
537: leads to activation law for conductivity with {\it linear} temperature
538: pre-exponent
539: \begin{equation}
540: \rho^{-1} \approx \sigma_0 \frac{BT}{W}\exp\left(-\frac{E_A}{T}\right),
541: \end{equation}
542: where $E_A=\Delta/2\approx 0.015W$ and $B$ is the following numerical constant
543: \begin{equation}
544: B=\lambda_c \sqrt{\frac{n_i}{n_c}}+ \lambda_i \sqrt{\frac{n_c}{n_i}}\approx 22
545: \end{equation}
546: for the parameters considered.
547:
548: Low values of conductivity obtained for the case of spin disorder
549: are
550: an indication of the possibility of Anderson localization \cite{kog,li,deph},
551: which
552: CPA is incapable of accessing. But the present results complement and support
553: the localization based approach. In fact, the results of Ref. \cite{kog}
554: were obtained under the assumption of the Fermi level pinning, which is now
555: explained as being due to strong electron-impurities interaction (and the
556: impurity band formation).
557:
558: In another aspect, the model considered may also explain low-temperature MIT observed in initially metallic manganites
559: R$_{1-x}$D$_x$MnO$_3$ upon substitution of R by isovalent atoms (e.g. La by Y \cite{barman}).
560: One may speculate that the substitution forms a deep impurity band which can capture holes
561: in R$_{1-x}$D$_x$MnO$_3$.
562:
563:
564: \section{Conclusion}
565: To conclude, we derived CPA equations for the one-electron Green function and
566: conductivity of
567: DE system containing impurities. The equations were solved for
568: the SC bare DOS and substitutional disorder model.
569: It was shown that
570: if the electron-impurity interaction is strong enough,
571: there is a gap between the
572: conduction band and the impurity band in PM phase,
573: the density of states being gapless in FM phase.
574: Under appropriate doping conditions the chemical potential is pinned inside the
575: gap. This can explain metal-insulator transition observed in manganites and
576: magnetic semiconductors.
577:
578: \section{Acknowledgments}
579:
580: This research was supported by the Israeli Science Foundation administered
581: by the Israel Academy of Sciences and Humanities.
582:
583: \begin{thebibliography}{99}
584:
585: \bibitem{helmolt} R. von Helmolt {\it et al}., Phys. Rev. Lett. {\bf 71}, 2331 (1993);
586: K. Chadra {\it et al}., Appl. Phys. Lett. {\bf 63}, 1990 (1993);
587: S. Jin {\it et al}., Science {\bf 264}, 413 (1994).
588:
589: \bibitem{ramirez} A.P. Ramirez, and M.A. Subramanian, Science, {\bf 277},
590: 546 (1998).
591:
592: \bibitem{sheng} L. Sheng, D.Y. Xing et. al, Phys. Rev. B {\bf 56}, R7053 (1997);
593: Phys. Rev. Lett. {\bf 79}, 1710 (1998).
594:
595: \bibitem{kog} E. M. Kogan and M. I. Auslender,
596: Phys. Stat. Sol. (b) {\bf 147}, 613 (1988).
597:
598: \bibitem{bebenin} N. G. Bebenin, R. I. Zainullina, V. V. Mashkautsan, V. S.
599: Gaviko, V. V. Ustinov, Y. M. Mukovskii, D. A. Shulyatev, JETP, {\bf 90},
600: 1027 (2000).
601:
602:
603: \bibitem{ziman} P. Soven, Phys. Rev. {\bf 156}, 809 (1967);
604: D. Taylor, Phys. Rev.
605: {\bf 156}, 1017 (1967);
606: J. M. Ziman, Models of Disorder, Cambridge University Press,
607: 1979.
608:
609: \bibitem{kubo} A. Rangette, A. Yanase, and J. Kubler, Solid State Comm. {\bf
610: 12}, 171 (1973); K. Kubo, J. Phys. Soc. Japan {\bf 36}, 32 (1974).
611:
612: \bibitem{takahashi} M. Takahashi, Phys. Rev. B {\bf 60}, 15858 (1999).
613:
614:
615: \bibitem{furukawa} N. Furukawa, J. Phys. Soc. Jpn. {\bf 64}, 2734 (1994);
616: N. Furukawa, cond-mat/9812066.
617:
618: \bibitem{moller} G. Moller, A. Ruckenstein, S. Schmitt-Rink, Phys. Rev. B {\bf
619: 46}, 7427 (1992); Th. Pruschke, D. L. Cox, M. Jarrel, Phys. Rev. B {\bf 47},
620: 3553 (1993).
621:
622: \bibitem{DMFA} A. Georges, G. Kotliar, W. Krauth, and M. J. Rozenberg, Rev. Mod.
623: Phys. {\bf 68}, 13 (1996).
624:
625: \bibitem{velic} B. Velicky, Phys. Rev. {\bf 184}, 614 (1969).
626:
627: \bibitem{khurana} A. Khurana, Phys. Rev. Lett. {\bf 64}, 1990 (1990).
628:
629: \bibitem{photo} M. v. Zimmermann, C.S. Nelson, J. P. Hill, Doon Gibbs, M.
630: Blume, D. Casa, B. Keimer, Y. Murakami, C.-C.
631: Kao, C. Venkataraman, T. Gog, Y. Tomioka, Y. Tokura, cond-mat/0007231.
632:
633: \bibitem{li} Quiming Li, Jun Zang and A.R. Bishop, C.M. Soukoulis,
634: Phys. Rev. B {56}, 4541 (1997).
635:
636:
637: \bibitem{deph} E. Kogan, M. Auslender, and M. Kaveh,
638: Eur. Phys. J. B {\bf 9}, 373 (1999).
639:
640: \bibitem{barman} A. Barman, M. Ghosh, S. Biswas, S.K. De and S. Chatterjee,
641: J. Phys. : Condens. Matter {\bf 10}, 9799 (1998).
642:
643: \end{thebibliography}
644:
645: \end{document}
646:
647:
648: