1: \documentclass[twocolumn,showpacs,amsmath,amssymb]{revtex4}
2: %\documentclass[preprint, showpacs,amsmath,amssymb]{revtex4}
3: \usepackage{graphicx}
4: \usepackage{bm}
5: \begin{document}
6: \title{Ferromagnetism from localized deep impurities in magnetic semiconductors.}
7: \author{Victor Barzykin}
8: \affiliation{Department of Physics and Astronomy, University of Tennessee,
9: Knoxville, TN 37996-1200}
10: \begin{abstract}
11: We propose that localized defects in magnetic semiconductors
12: act as deep impurities and can be described by the
13: Anderson model. Within this model, hybridization of
14: d-orbitals and p-orbitals gives rise to a non-RKKY indirect exchange mechanism,
15: when the localized d-electrons are exchanged through both conduction
16: and valence bands. For semiconductors with indirect band gap the non-RKKY part of exchange integral is
17: antiferromagnetic, which suppresses ferromagnetism.
18: In case of direct band gap, this exchange mechanism can, under certain conditions, lead to
19: enhancement of ferromagnetism.
20: The indirect exchange intergral is much stronger than RKKY, and can be sufficiently long range.
21: Thus, a potentially new class of high-temperature magnetic semiconductors emerges, where
22: doped carriers are not necessary to mediate ferromagnetism.
23: Curie temperatures in such magnetic semiconductors are determined mostly by the interaction between
24: localized impurities, \textit{not Zener mechanism}. This effect could also
25: be responsible for unusually high Curie temperatures in some magnetic semiconductors with
26: direct band gap, such as Ga$_{1-x}$Mn$_x$As.
27: \end{abstract}
28: \vspace{0.15cm}
29:
30: \pacs{75.50.Pp, 72.80.Ey, 75.30Hx, 75.50Dd}
31: \maketitle
32:
33: \section{Introduction}
34:
35: The advantage of ferromagnetic semiconductors (FS) as a source of
36: spin-polarized carriers is that they can be easily integrated into semiconductor devices\cite{Prinz, Wolf}.
37: When discovered, ferromagnetism at room temperatures with full polarization of itinerant carriers will be
38: a major breakthrough in semiconductor electronics. Most theoretical
39: and experimental efforts have been concentrated on III-V, group IV, and II-VI-based diluted magnetic semiconductors (DMS).
40: These semiconductors are alloys in which some atoms are randomly replaced by magnetic atoms, such as Mn$^{2+}$.
41:
42: Ferromagnetism in diluted magnetic semiconductors (DMS) is thought to be well understood in terms of the
43: so-called p-d exchange model,
44: which was first considered
45: over 50 years ago\cite{Zener1,Zener2,AG} (See also Ref.\cite{Korenblit, Furdyna, Dietll} for a review).
46: At concentrations of impurities above the Mott limit, i.e., as soon as carriers become delocalized,
47: conventional model of FS is fairly simple.
48: According to, for example, Ref.\cite{AG}, the interaction between
49: charge carriers in a semiconductor and spin-S impurities can be written as:
50: \begin{eqnarray}
51: \label{int}
52: U &=& - \int \bm{s}(\bm{r}) \sum_i \bm{S}_i J^{pd}(\bm{r} - \bm{R}_i) d^3 \bm{r} \\
53: &=& - J^{pd}_{\bm{q}=0} \bm{s} \sum_i \bm{S}_i, \nonumber
54: \end{eqnarray}
55: where $\bm{S}_i$ and $\bm{R}_i$ are the spin and the position of an i-th atom of magnetic impurity,
56: $J^{pd}_{\bm{q}=0} = \int J(\bm{r}) d^3 \bm{r}$, and we have used the fact that magnetic impurities are randomly distributed
57: in the sample. A simple analysis\cite{AG} then shows that, in the presence of one type of charge carriers, Curie temperature,
58: $T_c$, is proportional to concentration of magnetic impurities, $N_i$, and the square of the strength of the exchange
59: interaction, $a$:
60: \begin{equation}
61: T_c = {n_i S (S+1) (J^{pd}_{\bm{q}=0})^2 \chi_0 \over 12 \mu_0^2}
62: \label{Curie}
63: \end{equation}
64: Here $\mu_0$ is the magnetic moment of charge carriers, $\chi_0$ is the Pauli term in the spin susceptibility
65: in the absence of impurities. At sufficiently large carrier densities,
66: the spin polarization of the charge carriers at $T=0$ is given by:
67: \begin{equation}
68: s = {n_i J^{pd}_{\bm{q}=0} S \chi_0 \over 4 \mu_0^2}
69: \end{equation}
70: In general, for any carrier density, the spin polarization is given by Ze\`eman-split Fermi surface,
71: with the difference in chemical potentials for ``up'' and ``down'' spins given by:
72: \begin{equation}
73: \mu_{\uparrow} - \mu_{\downarrow} = J^{pd}_{\bm{q}=0} n_i S
74: \end{equation}
75: The spin polarization at any filling is then very easily calculated from this equation. For example
76: when the carrier density
77: \begin{equation}
78: n_e \le n_c = \frac{(2 J^{pd}_{\bm{q}=0} n_i S m^*)^{3/2}}{6 \pi^2}\, ,
79: \end{equation}
80: the carriers will be fully polarized at $T=0$. When $n_e > n_c$, they are no longer fully polarized.
81: The polarization of carriers is then determined by a parametric equation
82: (with $\mu$ as a parameter):
83: \begin{equation}
84: s = \frac{1}{2}\, (n_{e \uparrow} - n_{e \downarrow}),
85: \end{equation}
86: where
87: \begin{equation}
88: n_{e \uparrow, \downarrow} =
89: \frac{(2 m^* \mu \pm J^{pd}_{\bm{q}=0} n_i S m^*)^{3/2}}{6 \pi^2}\,,
90: \end{equation}
91: and $n_e = n_{e \downarrow} + n_{e \uparrow}$.
92:
93: Despite the simplicity of the basic concept of Zener ferromagnetism, calculations of $T_c$ for real materials
94: become rather involved, and depend crucially on details of the
95: band structure, the $p-d$ exchange matrix, or direct antiferromagnetic exchange between $Mn^{2+}$ spins.
96: Most theoretical and experimental efforts on magnetic semiconductors have been concentrated on
97: finding a new DMS-based material that would have a ferromagnetic transition above room temperature, and would be possible to
98: incorporate in thin film form with mainstream semiconductor device materials.
99: There are theoretical predictions for T$_c$'s above room temperatures in several classes of these materials\cite{dietl1,dietl2}.
100: However, experiments indicate that the growth of Curie temperature with concentration of magnetic impurities
101: saturates at 5-10\% Mn doping in most ferromagnetic semiconductors; it may even start to decrease at higher
102: concentrations of Mn. Since, according to the mean field p-d model, the Curie temperature Eq.(\ref{Curie})
103: grows linearly or faster (if the carrier concentration changes) with the growth of Mn concentration, it is
104: important to understand what limits the growth of $T_c$, and how this can be avoided.
105:
106: Experimental effort in FS concentrated mostly on In$_{1-x}$Mn$_x$As\cite{Munekata} ($T_c \sim 35K$)
107: and Ga$_{1-x}$Mn$_x$As\cite{deBoeck,Ohno,Ohno1,Ohno2} ($T_c \sim 110K$).
108: While room temperature ferromagnetism in FS remains a theoretical possibility, the highest Curie temperature reported in
109: the homogeneous sample with Mn concentration around 5\% is 116K for Mn$_x$Ge$_{1-x}$ \cite{Park}.
110: The saturation $T_c$ growth with increased Mn concentration is usually
111: ascribed to increased disorder. Disorder effectively introduces an exponential cutoff for the RKKY interaction and
112: reduces its range\cite{Ohno}. However, as we discuss in section VIII, weak disorder does not change Curie temperature.
113: Another possible reason is that
114: the strength of direct antiferromagnetic exchange grows as the average distance between impurities becomes shorter,
115: which, in turn, lowers the Curie temperature.
116: Room temperature ferromagnetism has been observed in a number of compounds with \textit{large} concentration of $Mn$
117: impurities (see Ref. \cite{Wolf} for a review), or, due to phase separation and formation of nanoclusters of these compounds (
118: such as Mn$_{11}$Ge$_8$ in Ref.\cite{Park}).
119: However, large carrier concentration in these compounds limits the degree of spin polarization that is necessary
120: for device applications.
121:
122: Some theoretical studies\cite{Millis} claimed that the reason Curie temperature can't get higher above
123: certain concentration of Mn is that the $p-d$ model
124: cannot be treated in the mean field approxination. However, since the RKKY interaction has a large radius,
125: mean field treatment in absence of strong disorder should be justified\cite{Larkin}.
126:
127:
128: In what follows, we adopt the idea that the reason for the discrepancy between mean field theory and experiment lies
129: in the presence of additional interactions in the effective spin Hamiltonian of the problem, which are not present
130: in the conventional formulation of the $p-d$ model. To derive effective spin Hamiltonian, including additional interactions,
131: we start from a more general Anderson model of deep magnetic impurities in FS.
132: We show that, in case of a deep Anderson impurity in a semiconductor,
133: an additional long-range indirect exchange interaction appears in the effective Hamiltonian, which, if
134: antiferromagnetic, severely limits $T_c$-s in these materials. On the other hand, for direct gap semiconductors, this
135: interaction could change sign and become ferromagnetic, if the magnetic impurity lies deep enough.
136: We derive conditions under which this becomes possible. As a result, ferromagnetic correlations would become
137: enhanced, not reduced, as the concentration of magnetic impurities grows, and high-temperature ferromagnetism could be
138: possible even without carriers.
139: The indirect exchange between two deep impurities, whether ferro- or antiferromagnetic, is stronger than the RKKY interaction,
140: and thus could produce large Curie temperatures. For example, it could provide a possible explanation of high-temperature
141: weak ferromagnetism\cite{Young} in La$_x$Ca$_{1-x}$B$_6$ or recently discovered
142: CaB$_2$C$_2$\cite{Akimitsu} (CaB$_2$C$_2$ has $T_c=770K$ and $M=10^{-4} \mu_B$).
143: Experiment\cite{Akimitsu,Fisk} indicates that these materails are direct band semiconductors with a relatively
144: small band gap, and that impurities play a major role in establishing the new high-temperature
145: ferromagnetic state. However, other reports\cite{Sato} claim that high-temperature ferromagnetism in these
146: materials is not a bulk effect. Rather, it is related to clustering
147: or new boron phases with Fe or Ni magnetic impurities. Understanding
148: the mechanism of weak ferromagnetism in these materials could lead to a discovery of more members of this class.
149: In either case, as we report in this paper, there is a theoretical
150: possibility of a new type of FS with direct band gap, where ferromagnetism is not carrier-driven.
151:
152: This paper is organized as follows. In Section II we discuss the general procedure of the derivation of the effective
153: low-energy Hamiltonian for Anderson impurities. In section III we discover that, due to strong hybridization,
154: an Anderson impurity is no longer a local center; its localized wave function acquires a finite range,
155: which is directly related to the corresponding "Bohr" radius of a charged impurity. Sections IV and V are devoted to
156: the derivation of the effective exchange Hamiltonian for magnetic impurities. In Section VI we
157: explore the consequences of the large range of interaction for magnetism, such as a high Curie temperature in case of a direct band
158: gap if magnetic impurities are dense enough. In section VII we consider a minor modification of the effective Hamiltonian in
159: case of higher spin (such as $S=5/2$ for Mn), and the application of these ideas to Ga$_{1-x}$Mn$_x$As.
160: In section VIII the influence of weak disorder and interactions on Curie temperature is briefly discussed.
161: Section XI provides a summary and conclusions.
162:
163: \section{The effective Hamiltonian}
164:
165: We start by considering a simple model of FS. In III-V systems, such as Ga$_{1-x}$Mn$_x$As,
166: it is well established that the Mn ions substitute for Ga, and contribute itinerant holes to the GaAs valence band. Experimentally,
167: the hole density is typically a small fraction (15\% or so) of the Mn concentration, perhaps due to strong localization of carriers
168: on Mn and other defects, so Ga$_{1-x}$Mn$_x$As can be considered partially compensated. The Mn ion has half-filled d-shell, which
169: acts like a spin-5/2 local moment. The Anderson model, which is more general
170: than the $p-d$ model, should completely account for all the physics of FS. It is well known that, when spins are well localized,
171: the single-impurity Anderson Hamiltonian is reduced to the $p-d$ Hamiltonian by the Schrieffer-Wolff transformation\cite{SW}.
172: For many impurities, this may no longer be the case. Let us start by considering a single-orbital Anderson Hamiltonian:
173:
174: \begin{equation}
175: H = H_0 + H_V,
176: \label{theH}
177: \end{equation}
178: where
179: \begin{eqnarray}
180: H_0 &=& \sum_{\bm{p} \sigma i} \epsilon_i(p) a^{\dagger}_{i \bm{p} \sigma} a_{i \bm{p} \sigma} + \\
181: &+& \sum_n \left[ \epsilon_0 \sum_{\sigma} d^{\dagger}_{n \sigma} d_{n \sigma} +
182: U d^{\dagger}_{n \uparrow} d^{\dagger}_{n \downarrow} d_{n \downarrow} d_{n \uparrow} \right]. \nonumber
183: \end{eqnarray}
184:
185: Here the first sum ($\bm{p}$) is taken over the reciprocal space, the second sum ($n$) over real space impurity sites.
186: $U$ is the on-site Coulomb repulsion term. Typically, $U$ is very large ($\sim 5 eV$), and can be taken to be infinite.
187: Here $\epsilon_i(p)$ are the energy band spectra of conduction and valence bands ($i=1,2$).
188:
189: The hybridization term in the model Hamiltonian, $H_V$, accounts for the $p-d$ hybridization between impurity sites and
190: conduction and valence bands:
191:
192: \begin{equation}
193: H_V = \frac{1}{N^{1/2}} \sum_{\bm{p} n \sigma i} V_i \left\{ a^{\dagger}_{\bm{p} \sigma i} d_{n \sigma} e^{-i \bm{p} \cdot \bm{R}_n}
194: + h.c.\right\}
195: \end{equation}
196:
197: This model is a reasonable generalization of the $p-d$ exchange Hamiltonian, usually considered in the literature.
198: Because of large on-site Coulomb
199: repulsion, the $d$-levels are half-filled. While we consider the case of a single d-orbital, a generalization
200: to $S=5/2$ Mn ion is straightforward (see section VII below).
201:
202: The Anderson Hamiltonian Eq.(\ref{theH}) describes a very complicated problem.
203: However, under the $U=\infty$ constraint, it can be reduced to the problem of Heisenberg spins
204: (in case of a single d-orbital, spin-$1/2$).
205: The low-energy effective
206: Hamiltonian is equivalent to Eq.(\ref{theH}) in the limit $k_B T \ll \Delta_i$, where $\Delta_i$ is the energy difference
207: between the impurity $d$-level and the top of the valence band, or the energy difference between the impurity $d$-level and the
208: bottom of conduction band. Since typical gap values in semiconductors are of the order of $1 eV$, this is usually a valid assumption.
209:
210: \begin{figure}
211: \includegraphics[width=3in]{fig1.eps}
212: %\includegraphics[width=3in]{fig1.pdf}
213: \caption{The second-order contribution to the effective Hamiltonian, shown as a time-ordered process.}
214: \label{fig2ord}
215: \end{figure}
216:
217: Following Refs. \cite{GS1,BG1}, the effective spin Hamiltonian can be derived by expanding the S-matrix
218: (or, at finite temperatures, the partition function) in $V$-s, and re-expressing various time-ordered processes
219: in terms of spin operators. Then, these processes are collected back under exponent, to obtain the effective
220: Hamiltonian. This method allows one to obtain
221: consistently the interaction between spins and carriers (electrons or holes) in conduction and valence bands, and
222: carrier-carrier interaction. In a way, the concept is similar to perturbative renormalization group, since
223: we arrive at a low energy effective Hamiltonian by integrating out higher-energy states. Treating hybridization term in Eq.(\ref{theH})
224: as a perturbation, we can rewrite the partition function as
225:
226: \begin{equation}
227: Z = Tr[\exp(-\beta H_0) S(\beta)],
228: \end{equation}
229: where
230: \begin{widetext}
231: \begin{equation}
232: S(\beta) = T \exp\left( - \int_0^{\beta} H_V(\tau) d \tau \right) =
233: T \exp(- \sum_{n \sigma i} \int_0^{\beta} d \tau V_i \left\{ \Psi^{\dagger}_{n \sigma i}(\tau) d_{n \sigma}(\tau)
234: + d^{\dagger}_{n \sigma}(\tau) \Psi_{n \sigma i}(\tau)\right\}
235: \label{smatr}
236: \end{equation}
237: \end{widetext}
238:
239: The problem of finding the effective Hamiltonian is then to reduce these expressions to the form:
240:
241: \begin{equation}
242: Z = Tr(\exp[-\beta H_{eff}]),
243: \end{equation}
244: using the $k_B T \ll \Delta_i$ condition, i.e., the fact that the local levels are almost always occupied, and that transitions to
245: conduction and valence bands are absent at low temperatures. Various terms in the effective Hamiltonian can then
246: be associated with certain time-ordered virtual processes.
247: For example, the first non-zero contribution to $H_{eff}$ is from the second order term in the expansion of
248: Eq.(\ref{smatr}) in $V_i$:
249: \begin{widetext}
250: \begin{equation}
251: S_2(\beta) = \sum_{n \sigma \sigma'} V_1^2 \int_0^{\beta} d \tau_1
252: \int_{\tau_1}^{\beta} d \tau_2 \Psi_{n \sigma 1}(\tau_2) \Psi^{\dagger}_{n \sigma' 1}(\tau_1)
253: d^{\dagger}_{n \sigma}(\tau_2) d_{n \sigma'} (\tau_1).
254: \label{2order}
255: \end{equation}
256: \end{widetext}
257: The order of operators $d^{\dagger}_{n \sigma}(\tau_2) d_{n \sigma'} (\tau_1)$ ($\tau_2 > \tau_1$) is fixed
258: by the assumption of strong Coulomb repulsion on the $n$-th center. Note that, in the second order, because
259: all centers are filled, only one band contributes to the effective Hamiltonian. The "filled" band does not contribute,
260: because of the Pauli principle (this is not the case at a finite U). The perturbative time-ordered process corresponding
261: to the term in Eq.(\ref{2order}) is shown in Fig.\ref{fig2ord}.
262:
263: We may rewrite $\Psi_{n \sigma 1}(\tau_2) \Psi^{\dagger}_{n \sigma' 1}(\tau_1)$ in the interaction representation
264: as:
265: \begin{widetext}
266: \begin{equation}
267: \Psi_{n \sigma 1}(\tau_2) \Psi^{\dagger}_{n \sigma' 1}(\tau_1) =
268: \{\Psi_{n \sigma 1}(\tau_2) \Psi^{\dagger}_{n \sigma' 1}(\tau_1)\}_+ -
269: \Psi^{\dagger}_{n \sigma' 1}(\tau_1) \Psi_{n \sigma 1}(\tau_2).
270: \label{commut}
271: \end{equation}
272: \end{widetext}
273: The first term in Eq.(\ref{commut}) is a c-number. In the limit $\Delta_1 \gg k_B T$ it is
274: possible to put $\tau_1 \simeq \tau_2$ in the second term in Eq.(\ref{commut}). Thus, for
275: the second-order contribution in the effective Hamiltonian we may write:
276: \begin{widetext}
277: \begin{equation}
278: H^{(2)}_{eff} = - \sum_{\bm{p}} \frac{V_1^2}{\epsilon_{1p} - \epsilon_0}\,
279: + \frac{V_1^2}{\Delta_1}\, \sum_{n \sigma \sigma'} \Psi^{\dagger}_{n \sigma' 1} \Psi_{n \sigma 1}
280: \left( \frac{1}{2}\, \delta_{\sigma \sigma'} + \bm{S}_n \bm{\sigma}_{\sigma' \sigma} \right),
281: \label{2oref}
282: \end{equation}
283: \end{widetext}
284: where $\bm{S}_n$ is the localized spin of the n-th impurity. The first term corresponds to the renormalization of
285: the energy of the localized level, while the second term involving spins of localized impurity
286: and carriers is nothing but the ordinary $H_{pd}$, the p-d model discussed in section I.
287:
288: The next order terms in the effective Hamiltonian are fourth order in V-s. There are sums
289: over two local centers, $m$ and $n$ in $S_4(\beta)$. Also, there are contributions to $S_4(\beta)$,
290: which we will denote $S'_4(\beta)$, that are already accounted for in the effective
291: Hamiltonian Eq.(\ref{2oref}):
292: \begin{equation}
293: S'_4(\beta) = \frac{1}{2}\, \int \int d \tau_1 d \tau_2 T\{H_{pd}(\tau_1) H_{pd}(\tau_2)\}.
294: \label{subtr}
295: \end{equation}
296: These contributions need to be subtracted from $S_4(\beta)$, to get the fourth-order (in $V_i$)
297: contributions in the effective Hamiltonian. The most important contribution
298: is the effective exchange interaction between localized spins:
299: \begin{equation}
300: H^{(4)}_{ex} = - \sum_{n \neq m} J(\bm{R}_n - \bm{R}_m) \bm{S}_n \bm{S}_m
301: \label{exchangeint}
302: \end{equation}
303: This exchange interaction is the result of 2 time-ordered processes shown in Figs. \ref{fig4orda}, \ref{fig4ordb}.
304: One is the superexchange (Fig. \ref{fig4orda}), which is a result of the localized spins exchanged
305: through the empty conduction band. The other process is the Bloembergen-Rowland term\cite{BR} (Fig. \ref{fig4ordb}),
306: an exchange process through both conduction and valence bands. The form of these contributions will be
307: discussed in greater detail in sections IV and V below.
308:
309: In addition, other interesting contributions arise as a result of fourth-order processes, such as p-d scattering by
310: spins on two centers,
311: \begin{widetext}
312: \begin{equation}
313: H^{(4)}_{pd2c} = \frac{V_1^2}{\Delta_1^2} \sum_{n \neq m, \sigma \sigma'} t(\bm{R}_n - \bm{R}_m)
314: \Psi^{\dagger}_{m \sigma' 1} \Psi_{n \sigma 1} [\bm{S}_n \bm{S}_m \delta_{\sigma' \sigma} +
315: i ([\bm{S}_m \bm{S}_n] \hat{\bm{\sigma}}_{\sigma' \sigma})],
316: \end{equation}
317: \end{widetext}
318: nontrivial local contribution,
319: \begin{widetext}
320: \begin{equation}
321: H^{(4)}_{local} = - \frac{V_1^4}{\Delta_1^3} \sum_{n \sigma \rho \rho'}
322: \Psi^{\dagger}_{n \sigma 1} \Psi^{\dagger}_{n \rho' 1} \left[\frac{1}{2}\, \delta_{\rho' \rho} +
323: (\hat{\bm{\sigma}}_{\rho' \rho} \bm{S}_n)\right] \Psi_{n \rho 1} \Psi_{n \sigma 1},
324: \end{equation}
325: \end{widetext}
326:
327: \begin{figure}
328: \includegraphics[width=3in]{fig2a.eps}
329: %\includegraphics[width=3in]{fig2a.pdf}
330: \caption{The fourth-order antiferromagnetic superexchange contribution to the effective exchange interaction between two localized
331: impurities.}
332: \label{fig4orda}
333: \end{figure}
334:
335: \begin{figure}
336: \includegraphics[width=3in]{fig2b.eps}
337: %\includegraphics[width=3in]{fig2b.pdf}
338: \caption{The fourth-order Bloembergen-Rowland term in the effective exchange interaction.}
339: \label{fig4ordb}
340: \end{figure}
341:
342: and corrections to the energy of the local level and ground state energy, which are dropped.
343: These nontrivial terms, however, are higher order in carrier density, which is small in magnetic
344: semiconductors. Thus, they don't play any significant role in magnetism, and, hence, will not be discussed in
345: any detail. We refer the reader to references \cite{GS1, GS2, BG1}, where these terms were discussed in detail
346: in connection with the 3-band model of cuprate superconductors.
347:
348: \section{Renormalization of local level due to hybridization.}
349:
350: To clarify the new physics of an Anderson impurity, in comparison with
351: $p-d$ magnetic impurity, let us first consider a single-level problem ($\bm{R}_n = 0$),
352: interacting only with the conduction band (i.e., we take $V_2=0$ for simplicity).
353: The Hamiltonian is defined on the manifold of wavefunctions $\{\Phi_0, \Psi_{1 \bm{p_1}},
354: \Psi_{1 \bm{p_2}}, ... \}$, which correspond to the
355: exact solution for a center in crystalline lattice in the absence of hybridization. These
356: wavefunctions form a complete orthogonal basis for the single-particle problem. We may now
357: introduce the hybridization as a perturbation into the Schr\"odinger equation. In the matrix representation,
358: the single-level Hamiltonian takes a very simple form:
359: \begin{equation}
360: H = \left\{ \begin{array}{cccc}
361: \epsilon_0 & v_1 & v_1 & \cdots \\
362: v_1 & \epsilon_{1 \bm{p}_1} & 0 & \cdots \\
363: v_1 & 0 & \epsilon_{1 \bm{p}_2} & \cdots \\
364: \vdots & \vdots & \vdots & \ddots
365: \end{array} \right\}
366: \end{equation}
367: Here $v_1 \equiv V_1/N^{1/2}$. This Hamiltonian can be
368: easily diagonalized, by looking for a solution for the ground state wave function as a linear
369: combination of all single-particle states:
370: \begin{equation}
371: \Psi = \chi_0 \Phi_0 + \sum_{\bm{p}} \{\chi^{(1)}_{\bm{p}} \Psi_{1 \bm{p}}\},
372: \end{equation}
373: where the coefficients obey the following set of equations:
374: \begin{eqnarray}
375: \label{spectrum}
376: (\epsilon_0 - \epsilon) \chi_0 + v_1 \sum_{\bm{p}} \chi_{\bm{p}} &=& 0 \\
377: v_1 \chi_0 + (\epsilon_{1 \bm{p}} - \epsilon) \chi_{\bm{p}}&=& 0 \nonumber
378: \end{eqnarray}
379:
380: \begin{figure}
381: \includegraphics[width=3in]{fig3.eps}
382: %\includegraphics[width=3in]{fig3.pdf}
383: \caption{Localized centers acquire a finite radius in the Anderson model. This radius can be as large
384: as 2-3 lattice spaces in magnetic semiconductors. An overlap of wavefunctions from two localized level
385: leads to the new physics, which is not captured by the traditional $p-d$ model.}
386: \label{fig3}
387: \end{figure}
388:
389: Solving Eq.(\ref{spectrum}) gives the new position of the localized level:
390: \begin{equation}
391: \epsilon = \epsilon_0 - \frac{V_1^2}{N}\, \sum_{\bm{p}} \frac{1}{\epsilon_{1 \bm{p}} - \epsilon}
392: \label{level}
393: \end{equation}
394: Thus, in presence of hybridization the position of the impurity level is shifted. If the depth
395: of the level $\Delta_1 \ll D_1$, where $D_1$ is the bandwidth of the conduction band (which
396: is usually a valid assumption for semiconductors, where the band gap is much smaller than the
397: corresponding bandwidths of conduction and valence bands), the sum in Eq.(\ref{level}) can
398: be performed numerically. The correction to the energy level,
399: \begin{equation}
400: \epsilon - \epsilon_0 \sim \frac{V_1^2}{D_1},
401: \end{equation}
402: turns out to be small for small enough
403: hybridization parameter $V_1 \ll \sqrt{D_1 \Delta_1}$. It is responsible for a
404: slight downward shift of the energy level. Such corrections will thus be dropped
405: in our further discussion.
406:
407: The wavefunction of the localized level, according to Eq.(\ref{spectrum}),
408: acquires an admixture to its decay in the form:
409: \begin{equation}
410: \delta \Phi_0(r) \sim V_1 m_1 a^2 \frac{a}{r}\, e^{-r \sqrt{2 m_1 \Delta_1}}.
411: \label{wavef}
412: \end{equation}
413: Here $a$ is the lattice constant (we assume a simple cubic lattice), $m_1$ is the effective mass
414: of carriers in the empty band (for hole-doped materials, the valence band is the ``empty'' band,
415: since it is empty of holes), $\Delta_1$ is the energy difference between impurity levels and the bottom
416: of the empty band.
417: Thus, a new length scale enters the problem,
418: \begin{equation}
419: R_0 = (2 m_1 \Delta_1)^{-1} \sim a \sqrt{D_1/\Delta_1},
420: \label{scale}
421: \end{equation}
422: which, while small compared to the average distance between doped carriers, may
423: considerably exceed interatomic distances (the large parameter is the square root of
424: the ratio of the bandwidth to the energy gap in a semiconductor), and become comparable to the average
425: distance between magnetic impurities. The overlap of the localized wave functions
426: gives rise to a new contribution to the exchange integral, which will be derived
427: below. The picture of finite-radius localized magnetic impurities with overlapping wavefunctions is
428: shown in Fig.\ref{fig3}.
429:
430: Existence of a large (compared to lattice spacing) radius for deep impurities was first noticed by
431: Keldysh\cite{Keldysh}, who analyzed deep charged impurities in ordinary semiconductors.
432: In his case, however, this scale corresponded to ordinary Bohr radius for a deep impurity
433: ( modified by the fact that in that case one had to consider Dirac Hamiltonian for the
434: $k-p$ model of semiconductors). In our case, the large length scale comes from
435: the hybridization of the localized impurity level with conduction and valence bands.
436: We will also see below in Section VI that effective length scale, at which direct interaction between impurities
437: starts to matter, grows logarithmically with decreasing temperature in the framework of percolation theory or
438: virial expansion. Thus, at low temperatures, such as Curie temperature, the effective range of interaction
439: between two impurities becomes even larger.
440:
441: \section{The exchange Hamiltonian.}
442:
443: \begin{figure}
444: \includegraphics[width=3in]{fig4.eps}
445: %\includegraphics[width=3in]{fig4.pdf}
446: \caption{Positions of conduction and valence bands.}
447: \label{bandposition}
448: \end{figure}
449:
450: The problem of magnetism in the Anderson model can be reduced to the spin Hamiltonian
451: Eq.(\ref{exchangeint}), described by the processes shown in Figs \ref{fig4orda}, \ref{fig4ordb}. In case
452: when the empty band (the "empty" band can also be the valence band, in case of hole doping)
453: gets partially filled by carriers (electrons or holes), the time-ordered
454: processes shown in Figs \ref{fig4orda}, \ref{fig4ordb}
455: include all possible contributions - the familiar RKKY exchange interaction, the superexchange,
456: and the Bloembergen-Rowland interaction. In presence of carriers, the superexchange gets modified
457: by the RKKY exchange term. We shall see below, that the Bloembergen-Rowland interaction also gets
458: modified in presence of dopants. This modification is not accounted for by the
459: RKKY interaction. As we have discussed in the previous section, localized impurities in the
460: Anderson model acquire a large radius (compared with the lattice constant, $a$). Thus, exchange
461: interaction between localized spins is, in general, long range. This would allow
462: one to use mean field to describe the ferromagnetic ordering, if the concentration of localized spins
463: is high enough. On the other hand, if the concentration of magnetic impurities is low, one can
464: use percolation theory or virial expansion.
465:
466: Let us first consider exchange integrals given by the processes shown in
467: Figs \ref{fig4orda}, \ref{fig4ordb} for a semiconductor with conduction and valence
468: bands separated by some general reciprocal space wavevector $\bm{Q}$, as shown in Fig.\ref{bandposition}.
469: When $\bm{Q}=0$, it is a direct band gap semiconductor.
470: Surprisingly, all exchange processes can be written in a relatively compact way:
471: \begin{equation}
472: J(R) = \sum_{i,j} J_{ij}(R),
473: \label{exchch}
474: \end{equation}
475: where
476: \begin{widetext}
477: \begin{equation}
478: J_{ij} = 2 \frac{\left|V_i\right|^2 \left|V_j\right|^2 a^6}{(2 \pi)^6}\,
479: \int d^3 \bm{p} d^3 \bm{q} \frac{[1 - n(\epsilon_{\bm{p} i} - \mu)] e^{i \bm{q} \bm{R}}}{
480: (\epsilon_{\bm{p} i} - \epsilon_{\bm{p}+\bm{q} j})(\epsilon_{\bm{p} i} - \epsilon_0)^2}.
481: \label{exch}
482: \end{equation}
483: \end{widetext}
484:
485: \begin{figure}
486: \includegraphics[width=3in]{fig5.eps}
487: %\includegraphics[width=3in]{fig5.pdf}
488: \caption{Superexchange contribution to exchange integral in q-space; modification by RKKY in presence of carriers
489: is shown by a dotted line.}
490: \label{superexchange}
491: \end{figure}
492: \begin{figure}
493: \includegraphics[width=3in]{fig6.eps}
494: %\includegraphics[width=3in]{fig6.pdf}
495: \caption{Bloembergen-Rowland contribution to exchange integral in q-space; modification in presence of carriers
496: is shown by a dotted line.}
497: \label{BlR}
498: \end{figure}
499:
500: Here ${i,j}$ are the band indices for conduction and valence bands, $\mu$ is the chemical potential,
501: $n(\epsilon - \mu)$ is the Fermi-Dirac distribution function. The empty band corresponds to $i=1$,
502: while the filled band has the band index $i=2$. We can see now, that for $T \ll \Delta_0$, where $\Delta_0$ is the band gap,
503: the contributions $J_{21}$ and $J_{22}$ are absent. $J_{11}$ is the superexchange, which corresponds to
504: the process shown in Fig. \ref{fig4orda}. In presence of carriers, this process, and the corresponding
505: expression, also includes the RKKY contribution. When carriers are present, the
506: effective Hamiltonian can be written in terms of impurity spins only, with carriers "integrated out".
507: In this case, impurity spins interact with the exchange Hamiltonian Eqs.(\ref{exchch}),(\ref{exch}), which
508: includes RKKY. If, however, we wish to retain carriers, and the carrier-impurity p-d exchange interaction,
509: as we described in section II, then the RKKY part will be the part $S'(\beta)$ in the 4-th order
510: (see Eq.(\ref{subtr})), which has to be subtracted, and the indirect exchange interaction will be
511: described by Eq.(\ref{exch}) at zero doping, i.e., with the factor $[1 - n(\epsilon_{\bm{p} i} - \mu)]$ in
512: Eq.(\ref{exch}) replaced by $1$. These two descriptions are only equivalent when the chemical
513: potential for doped carriers, $\mu \ll \Delta_1$. Basically, when the doped carriers are retained, we "integrate out" electrons
514: and holes in the original Anderson Hamiltonian up to the scale $\mu$. When we consider the Hamiltonian
515: for localized spins only, we integrate out all carriers.
516: It makes sense to define the RKKY interaction
517: for the Anderson model as the difference between the doped and undoped cases in Eqs.(\ref{exchch}),(\ref{exch}).
518: When $\mu \sim \Delta_1$, the Anderson model \textit{cannot} be reduced to the $p-d$ model, except near the Fermi surface.
519: The Anderson model expression for the RKKY interaction in real space
520: will then \textit{differ} from the corresponding $p-d$ model expression
521: at \textit{short} distances, which are important for ferromagnetism. We won't consider
522: this subtle point any further, since we will always assume the carrier density in FS to be small,
523: i.e., $\mu \ll \Delta_1$. In this section and the next section we will deal with an
524: effective Hamiltonian for spins only, i.e., when all carriers are integrated out. This procedure
525: for the Anderson model is always justified (although, as we commented above, a reduction to the $p-d$ model is \textit{not}).
526: $J_{12}$ is the Bloembergen-Rowland interaction, which
527: also gets somewhat modified by doped carriers. Note that the expression for the exchange integral Eq.(\ref{exch}),
528: obtained from the Anderson model, differs from the original expression of Bloembergen and Rowland \cite{BR};
529: it includes the term $(\epsilon_{\bm{p} 1} - \epsilon_0)^2$ in the denominator, which makes the
530: exchange integral in the Anderson model long range. Both contributions to exchange integral can be evaluated
531: numerically and analytically; they are shown schematically in Figs \ref{superexchange} and \ref{BlR}.
532:
533: We see that the superexchange is, in general, antiferromagnetic, while the Bloembergen-Rowland
534: contribution favors a spin density wave with a wavevector $\bm{Q}$, separating conduction and valence bands.
535: Both exchange processes are, in general, of the same order, and their relative strength is determined by
536: the corresponding hybridization parameters $V_1$ and $V_2$. The superexchange integral in the $\bm{q}$ - space,
537: in the absence of carriers, can be easily found analytically:
538: \begin{equation}
539: J_{11}(\bm{q}) = - \frac{2 V_1^4 a^3 m_1^3}{\pi \sqrt{2 m_1 \Delta_1}}\, \frac{1}{q^2 + 8 m_1 \Delta_1}
540: \label{superex}
541: \end{equation}
542: In presence of carriers, $J_{11}(\bm{q})$ gets modifed by the RKKY interaction:
543: \begin{widetext}
544: \begin{equation}
545: \delta J_{11}(\bm{q}) \equiv J_{RKKY}(\bm{q}) = \frac{V_1^4 a^3 m_1 p_F}{(2 \pi)^2 \Delta_1^2} \,
546: \left\{ 1 + \left(\frac{p_F}{q}\, - \frac{q}{4 p_F}\, \right) \ln
547: \left(\frac{1 + \frac{q}{2 p_F}\,}{1 - \frac{q}{2 p_F}\,}\right)\right\},
548: \end{equation}
549: \end{widetext}
550: where $p_F = \sqrt{2 m_1 \mu}$ is the Fermi momentum for doped carriers. The superexchange and RKKY contributions
551: can be easily rewritten in real space:
552: \begin{equation}
553: J_{11}(R) = J_{se}(R) + J_{RKKY}(R),
554: \label{j11}
555: \end{equation}
556: with
557: \begin{equation}
558: J_{se}(R) = - \frac{V_1^4 a^6 m_1^3}{2 \pi^2 \sqrt{2 m_1 \Delta_1} R}\, \exp\{-\sqrt{8 m_1 \Delta_1} R\},
559: \label{sesese}
560: \end{equation}
561: and a standard expression for the RKKY interaction, taking into account that $J_{pd} \equiv 2 V_1^2/\Delta_1$
562: (note that in Eq.(\ref{exchangeint}) we sum over each pair of impurities twice):
563: \begin{equation}
564: J_{RKKY}(R) = - J_{RKKY} F(2 p_F R),
565: \end{equation}
566: where
567: \begin{equation}
568: F(x) = \frac{\cos(x)}{x^3}\, - \frac{\sin{x}}{x^4},
569: \end{equation}
570: and
571: \begin{equation}
572: J_{RKKY} = \frac{V_1^4 a^6 m_1 p_F^4}{\pi^3 \Delta_1^2}.
573: \end{equation}
574: We define the RKKY interaction in the Anderson model formulation as the difference between $J_{11}(R)$
575: in doped and undoped cases, given by Eq.(\ref{exch}). As we emphasized above, since it depends explicitly
576: on impurity energy level, when $\Delta_1 \sim \mu$, this expression differs from the usual RKKY form.
577: Its asymptotic behavior at large distances, certainly, does not change.
578: In what follows we assume that the carrier concentration is always low, $\mu \ll \Delta_i$, so that
579: $J_{RKKY}(R)$ is given by the standard expression.
580: Note that if $J_{pd}$ and the level position are known, the exchange integral Eq.(\ref{j11}) contains
581: no free parameters:
582: \begin{widetext}
583: \begin{equation}
584: J_{11}(R) = - J_{RKKY}
585: \left(F(2 p_F R) + \frac{\pi (m_1 \Delta_1)^{3/2}}{2 \sqrt{2} p_F^4 R}\, e^{-2 \sqrt{2 m_1 \Delta_1} R} \right).
586: \label{11ex}
587: \end{equation}
588: \end{widetext}
589: Similar to superexchange, the Bloembergen-Rowland interaction can also be written in two parts. The first
590: part is from the empty and filled bands, while the second part takes doped carriers into account:
591: \begin{equation}
592: J_{12}(\bm{q}) = J_{BR}(\bm{q}) + J'_{12}(\bm{q}).
593: \label{12ex}
594: \end{equation}
595: The first contribution in Eq.(\ref{12ex}) favors a spin density wave with a wavevector $\bm{Q}$, separating
596: conduction and valence bands, and has a rather cumbersome form:
597: \begin{widetext}
598: \begin{equation}
599: J_{BR}(\bm{q}) = \frac{a^3 V_1^2 V_2^2 m_1^2 m_2}{\pi \sqrt{2 m_1 \Delta_1}}\,
600: \left\{\frac{m_1 \Delta_1 + m_2 \Delta_1 + \frac{q^2}{2}\, + m_2 \Delta_0 - \sqrt{2 (m_1 + m_2) \Delta_1}
601: \sqrt{2 m_2 \Delta_0 + q^2 \frac{m_2}{m_1 + m_2}\,}}{\Delta_1^2(m_1+m_2)^2 + \Delta_1 (q^2 (m_1-m_2) -
602: 2 m_2 (m_1+m_2) \Delta_0) + \left(\frac{q^2}{2}\,+ m_2 \Delta_0 \right)^2}\right\}.
603: \label{BRl}
604: \end{equation}
605: \end{widetext}
606: Here and below in this section the notation will be slightly different; $\bm{q}$ is now the \textit{difference}
607: from $\bm{Q}$ (the wavevector separating the bottoms of the two bands), i.e., we
608: change our notation in the following way: $\bm{q - Q} \longrightarrow \bm{q}$.
609: In general, the Bloembergen-Rowland contribution Eq.(\ref{BRl}) is quite cumbersome, which makes it impossible to
610: find an explicit analytical expression
611: in real space, except in some limiting cases. However, the leading asymptotic behavior of $J_{BR}(R)$ at large distances
612: can be derived. If the deep impurity
613: level lies inside the gap, $\Delta_1 < \Delta_0$, we find an exponential decay of exchange correlations:
614: \begin{widetext}
615: \begin{equation}
616: J_{BR}(\bm{R}) \simeq \frac{a^6 V_1^2 V_2^2 m_1 m_2}{2 \sqrt{2} \pi^2 R}\,\left(\sqrt{\frac{m_1}{\Delta_1}\,} -
617: \sqrt{\frac{m_2}{\Delta_0 - \Delta_1}\,} \right) \cos{(\bm{Q} \cdot \bm{R})} e^{-R/R_1},
618: \label{undBR}
619: \end{equation}
620: \end{widetext}
621: where the range of the integral is:
622: \begin{equation}
623: R_1 = \frac{1}{\sqrt{2 m_1 \Delta_1} + \sqrt{2 m_2 (\Delta_0 - \Delta_1)}}.
624: \label{r1}
625: \end{equation}
626: On the other hand, when the deep impurity level enters the filled band, $\Delta_1 > \Delta_0$, the Bloembergen-Rowland
627: contribution has an oscillating decaying asymptotic:
628: \begin{widetext}
629: \begin{equation}
630: J_{BR}(\bm{R}) = \frac{a^6 V_1^2 V_2^2 m_1 m_2}{ \sqrt{2} \pi^2 R}\, \left(\sqrt{\frac{m_1}{2 \Delta_1}\,} \cos\frac{R}{R_{c1}}\, +
631: \sqrt{\frac{m_2}{2 (\Delta_1-\Delta_0)}\,} \sin\frac{R}{R_{c1}}\,\right) \cos{(\bm{Q} \cdot \bm{R})} e^{- R/R_{c0}},
632: \end{equation}
633: \end{widetext}
634: where
635: \begin{equation}
636: R_{c1} = \frac{1}{\sqrt{2 m_1 (\Delta_1 - \Delta_0)}}, \ \ \ R_{c0} = \frac{1}{\sqrt{2 m_1 \Delta_1}}\, .
637: \label{rc1rc0}
638: \end{equation}
639:
640: The contribution from doped carriers to $J_{12}$ can be derived as well:
641: \begin{equation}
642: J'_{12}(\bm{q}) \simeq - \frac{ a^3 V_1^2 V_2^2 p_F^3}{3 \pi^2 \Delta_1^2 (\Delta_0 + \frac{q^2}{2 m_2}\,)}\,
643: \end{equation}
644: In real space, this contribution becomes:
645: \begin{equation}
646: J'_{12}(\bm{R}) \simeq - \frac{p_F^3 a^6 V_1^2 V_2^2 m_2}{6 \pi^3 \Delta_1^2 R} \cos{(\bm{Q} \cdot \bm{R})} e^{-\sqrt{8 m_2 \Delta_0} R}
647: \label{dopBR}
648: \end{equation}
649:
650: We see that if $\bm{Q}=0$, i.e., when our FS has a direct band gap, the Bloembergen-Rowland mechanism gives
651: a large ferromagnetic short-range contribution to the exchange integral.
652: At high enough concentration of impurities, ferromagnetic properties are then determimed by the value of exchange
653: integral at $\bm{q}=0$. This value can be easily written down (it is only valid for a direct band gap FS):
654:
655: \begin{widetext}
656: \begin{equation}
657: J(\bm{q}=0) = J_{11}(\bm{q}=0) + J_{12}(\bm{q}=0)
658: = \frac{2 a^3 V_1^2 m_1^2}{ \pi \sqrt{2 m_1 \Delta_1}}\, \left(- \frac{V_1^2}{8 \Delta_1}\,
659: + \frac{V_2^2 m_2}{ (\sqrt{2 (m_1 + m_2) \Delta_1} + \sqrt{ 2 m_2 \Delta_0})^2}\, \right)
660: \label{qzero}
661: \end{equation}
662: \end{widetext}
663:
664: In this section we have derived explicitly the interaction between two magnetic impurities in a semiconductor
665: with one conduction and one valence band. Typically, the band structure of semiconductors is more complex
666: than that (for example, GaAs has light and heavy hole bands).
667: This situation is considered in the next section.
668:
669:
670: \section{What if a semiconductor has more than two bands?}
671:
672: When a semiconductor has many bands, the analysis is just as straightforward as in case
673: of two bands considered in the previous section. The exchange integral between two impurities
674: is still given by Eqs(\ref{exchch}),(\ref{exch}), where the sum now goes over all pairs of band indices. The case when
675: $i=j$ is an ``empty'' band index corresponds to a superexchange contribution
676: (Fig.\ref{fig4orda}), considered in detail in the previous section; $J_{ij} = 0$ when $i$ is a ``filled''
677: band index, because of the Pauli principle; $J_{ij}$ is the Bloembergen-Rowland contribution (Fig.\ref{fig4ordb}),
678: when $i$ is an ``empty''
679: band index, and $j$ is a ``filled'' band index. The contribution of a new type appears when one has
680: two or more ``empty'' bands in a semiconductor. Then there will be a superexchange contribution of the type shown in
681: Fig.\ref{fig4orda}, where now the carriers are exchanged through two different ``empty'' bands - the process (1) in
682: Fig.\ref{fig4orda} puts a carrier from one magnetic impurity into the first ``empty'' band, while the process (2)
683: puts a carrier from the other magnetic impurity into the second ``empty'' band. Naturally,
684: there are two such contributions, $J_{ij}$ and $J_{ji}$, where $i \neq j$ indices correspond to two
685: different ``empty'' bands. The exchange Hamiltonian in any FS
686: is a sum of all pairwise contributions listed above.
687:
688: Let us now consider the new two-band superexchange contribution in detail.
689: In the most general case,
690: the bottom of the second ``empty'' band is shifted from the bottom of the first ``empty'' band by a
691: wavevector $\bm{Q}$. The bottoms of the two bands also lie at two different positive
692: energies $\Delta_1$ and $\Delta_2$ relative to impurity level, unless there is a symmetry-related
693: degeneracy, and the carriers in the first and second bands have different
694: effective masses $m_1$ and $m_2$.
695:
696: Let us first consider the simplified case when the bands 1 and 2 are identical (i.e., $\Delta_1 = \Delta_2$ and
697: $m_1 = m_2$), but separated by a wavevector $\bm{Q}$. Then we don't have to do the calculation, since
698: $\epsilon_{\bm{p}2} = \epsilon_{\bm{p-Q} 1}$, and it can
699: be easily seen from Eq.(\ref{exch}), that
700: \begin{equation}
701: J_{12}(\bm{q}) = J_{11}(\bm{q-Q}),
702: \end{equation}
703: and
704: \begin{equation}
705: J_{12}(R) = J_{11}(R) e^{i \bm{Q} \bm{R}}.
706: \end{equation}
707: Summing all contributions from 2 identical empty bands, since $J_{11}(R) = J_{22}(R)$,
708: we get:
709: \begin{equation}
710: J_{2bse} = 2 J_{11}(R) [1 + \cos{(\bm{Q} \cdot \bm{R})}].
711: \label{2bse}
712: \end{equation}
713: This also includes RKKY-type contribution for two identical bands at finite doping, described by:
714: \begin{equation}
715: J_{12 RKKY} + J_{21 RKKY} = 2 J_{RKKY}(R) \cos{(\bm{Q} \cdot \bm{R})},
716: \label{2bse1}
717: \end{equation}
718: and the ordinary RKKY interaction from both bands.
719: It is rather obvious from Eq.(\ref{2bse1}), that at a finite density of carriers there are ordinary one-band RKKY contributions from
720: both bands. The two-band RKKY contribution oscilates much more rapidly in space,
721: as $\cos(\bm{Q}{R}))$, unless $\bm{Q = 0}$. In the latter case, the two-band RKKY contribution just enhances
722: the contributions from the two separate bands. In the case when $\bm{Q}=0$, the $12$ and $21$ exchange integrals take
723: the same form as shown in Fig.\ref{superexchange}. For $\bm{Q \neq 0}$, these contributions have the
724: same form in $\bm{q}$-space as for $\bm{Q}=0$, centered at the wavevector $\bm{q}=\bm{Q}$.
725:
726: Now that the form of new two-band contributions to the exchange integral has become clear from the simplified case,
727: let us consider a more
728: general case of two completely different ``empty'' bands. The bottom of the second band is shifted from
729: the bottom of the first band by the wavevector $\bm{Q}$. Of course, $12$ and $21$ contributions to $J(\bm{q})$
730: will still be of the form shown in Fig.\ref{superexchange}, with the bottom at $\bm{q} = \bm{Q}$, but the corresponding
731: expressions become more combersome. Let us assume that $\Delta_1 < \Delta_2$, i.e. the bottom of the first band
732: lies below the bottom of the second band, and the first band gets filled
733: by carriers first. We consider three separate cases, (1) both bands are empty; (2) the first band gets
734: filled by carriers, but the second band is empty, and (3) both bands get partially filled by carriers.
735: When $\Delta_1 = \Delta_2$, i.e., the two bands are symmetry-related, we only have cases (1) and (3).
736: As before, we assume that the carrier
737: concentration is very small, i.e., the Fermi energy for doped carriers is much smaller than any other
738: energy scale in the problem, except temperature.
739:
740: \vglue 0.5cm
741:
742: \noindent
743: \textit{(1) Both bands are empty.}
744:
745: \vglue 0.5cm
746:
747: Let us now consider the exchange interaction arising from two empty bands. The integral
748: in Eq.(\ref{exch}) can be easily calculated:
749: \begin{widetext}
750: \begin{equation}
751: J_{12+21}(\bm{q}) = - \frac{\sqrt{2} V_1^2 V_2^2 m_1 m_2 a^3}{(\bm{q}-\bm{Q})^2 +
752: (\sqrt{2 m_1 \Delta_1} + \sqrt{2 m_2 \Delta_2})^2}\, \left(\sqrt{\frac{m_1}{\Delta_1}\,}
753: +\sqrt{\frac{m_2}{\Delta_2}\,}\right).
754: \end{equation}
755: \end{widetext}
756: In the coordinate space, it takes the following form:
757: \begin{equation}
758: J_{12+21}(\bm{R}) = - V_{12} \cos{(\bm{Q}\cdot\bm{R})} \frac{1}{R} \exp{( - \frac{R}{R_{12}}\, )},
759: \label{2eb}
760: \end{equation}
761: where the range of this interaction is given by:
762: \begin{equation}
763: R_{12} = \frac{1}{\sqrt{2 m_1 \Delta_1} + \sqrt{2 m_2 \Delta_2}},
764: \end{equation}
765: and
766: \begin{equation}
767: V_{12} = \frac{V_1^2 V_2^2 a^6 m_1 m_2}{2 \sqrt{2} \pi^2}\, \left(\sqrt{\frac{m_1}{\Delta_1}\,} +
768: \sqrt{\frac{m_2}{\Delta_2}\,} \right)
769: \end{equation}
770:
771: \vglue 0.5cm
772:
773: \noindent
774: \textit{(2) One band gets filled.}
775:
776: \vglue 0.5cm
777:
778: Let us assume in this section that band one gets filled first, i.e., $\Delta_1 < \Delta_2$.
779: We also assume that the chemical potential (counted from the bottom of band one), $\mu_1 \ll \Delta_2 - \Delta_1$,
780: i.e., the number of carriers is low. Of course, the main part of the exchange inegral is still given by Eq.(\ref{2eb}).
781: However, at finite doping there are corrections. Obviously, since only band one gets carriers, $\delta J_{21} = 0$.
782: At small filling, RKKY-like correction can be easily calculated:
783: \begin{equation}
784: \delta J_{12}(q) \simeq \frac{2 V_1^2 V_2^2 a^3 P_{F1}^3 m_2}{3 \pi^2 \Delta_1^2 (2 m_2 [\Delta_2 - \Delta_1] + q^2)}.
785: \end{equation}
786: Here $q$ denotes the difference from $\bm{Q}$, the wavevector separating the bottoms of the two bands.
787: In coordinate space it can be written as:
788: \begin{equation}
789: \delta J_{12}(\bm{R}) \simeq
790: \frac{V_1^2 V_2^2 a^6 p_{F1}^3 m_2}{6 \pi^3 \Delta_1^2 R} e^{- R/R_g} \cos{\bm{Q} \cdot \bm{R}},
791: \end{equation}
792: where
793: \begin{equation}
794: R_g = \frac{1}{\sqrt{2 m_2 (\Delta_2 - \Delta_1)}}\,
795: \end{equation}
796:
797: \vglue 0.5cm
798:
799: \noindent
800: \textit{(3) Both bands get filled.}
801:
802: \vglue 0.5cm
803:
804: When $\Delta_1 = \Delta_2 = \Delta$ by symmetry, both bands get filled by carriers simultaneously.
805: The result of this is an RKKY-like correction to exchange integral Eq.(\ref{2eb}), which is given
806: by Eq.(\ref{2bse1}) for two identical bands. If the bands have different masses, the interaction
807: is RKKY-like long range, but the general expression is quite messy. Here we give the its value at
808: $\bm{q}=\bm{Q}$, which is important for ferromagnetism, when $\bm{Q} = 0$:
809:
810: \begin{equation}
811: \delta J_{12+21}(\bm{q}=\bm{Q}) = \frac{2 V_1^2 V_2^2 a^3 \sqrt{2 \mu} m_1 m_2}{\pi^2 \Delta^2 (\sqrt{m_1} + \sqrt{m_2})}\,
812: \end{equation}
813:
814: To summarize this section, the exchange interaction in any magnetic semiconductor is composed of pairwise contributions:
815: the superexchange contribution through one ``empty'' band, the Bloembergen-Rowland contribution from one ``empty'' and one filled band,
816: and a pairwise contribution from two different ``empty'' bands. The last contribution was considered in this section.
817: It includes two-band long-range RKKY-like interaction at finite doping, which is usually not taken into account in
818: the $p-d$ Hamiltonian. Note that when $\bm{Q} \neq 0$, the two-band RKKY interaction will favor spin glass order,
819: \textit{not} ferromagnetism, since it oscillates very rapidly in real space.
820:
821: \section{The mean field approximation, virial expansion, and percolation.}
822:
823: Now that we have obtained all terms in the exchange interaction Eq.(\ref{exch}), we can proceed to
824: calculate magnetic properties of ferromagnetic semiconductors. In general, the RKKY exchange interaction
825: is always long-range. The shorter-range contributions may or may not be treated in mean field, depending
826: on the ratio of their range to the average inter-impurity distance. We assume here, for simplicity,
827: that one of indirect exchange processes considered in the previous two sections dominates the short-range
828: physics. In case of FS with indirect band gap, the Bloembergen-Rowland exchange contribution favors
829: antiferromagnetism with a lattice wavevector $\bm{Q}$. It oscillates very rapidly in real space.
830: Since impurities are distributed randomly, it would favor a spin glass ordering.
831: The contributions at $\bm{Q}=0$, which influence ferromagnetism, are the RKKY
832: and the superexchange. From Eq.(\ref{11ex}) one can easily see that, when impurity concentration
833: $n_i \gg (m_1 \Delta_1)^{3/2}$, the superexchange contribution totally suppresses ferromagnetism
834: brought about by the
835: RKKY exchange interaction. On the other hand, when $n_i \ll (m_1 \Delta_1)^{3/2}$, the superexchange
836: at an average distance between impurities is
837: suppressed, and the RKKY interaction gives rise to ferromagnetism, as in the ordinary $pd$-model. Corrections
838: due to superexchange can then be calculated using the virial expansion approach\cite{BA}. We can
839: rewrite the dominant exchange contribution from Eq.(\ref{11ex}) in the following form:
840: \begin{equation}
841: J(R) = - J_{RKKY} F(2 p_F R) - V_0 \frac{R_0}{R}\, e^{-R/R_0},
842: \label{generic}
843: \end{equation}
844: where $V_0 = \pi J_{RKKY}/(64 p_F^4 R_0^4)$, $R_0 = 1/\sqrt{8 m_1 \Delta_1}$, in case when it is given by one empty band only.
845: If contributions from more than one ``empty'' band are important, $V_0$ and $J_{RKKY}$ are, in general, unrelated to each
846: other. The reason for this is that the largest short range contribution typically comes from the lightest bands,
847: which would produce the longest range indirect exchange (if all hybridization parameters
848: are of the same order). On the other hand, the dominant
849: RKKY contribution is the one from the heaviest band, since the carriers in that band have the largest Fermi wavevector; there
850: will also be additional RKKY contributions from many bands, which were analyzed in the previous section.
851: In what follows we consider the simplest case, when only one heavy band is relevant for RKKY.
852:
853: In case of a FS with a direct band gap, the Bloembergen-Rowland mechanism gives rise to ferromagnetism, and
854: we can either have a ferromagnetic or an antiferromagnetic short-range exchange contribution, depending
855: on the relative strength of corresponding exchange processes. If the Boembergen-Rowland term
856: dominates the physics at short distances, the impurity spin Hamiltonian becomes:
857: \begin{equation}
858: J(R) = - J_{RKKY} F(2 p_F R) + V_{BR} \frac{R_1}{R}\, e^{-R/R_1},
859: \end{equation}
860: In general, the Bloembergen-Rowland term Eq.(\ref{12ex}) also has an antiferromagnetic contribution from
861: doped carriers. It then can be rewritten in the following form:
862: \begin{equation}
863: J_{12}(R) = V_{BR} \frac{R_1}{R} e^{-R/R_1} - V_{dop} \frac{R_g}{R_0} e^{-R/R_g},
864: \end{equation}
865: where $R_g = 1/\sqrt{8 m_2 \Delta_0}$, $V_{BR}$ and $V_{dop}$ are given by the coefficients
866: in front of exponents in Eq.(\ref{undBR}) and Eq.(\ref{dopBR}). However, since
867: \begin{equation}
868: V_{dop}/V_{BR} \sim \frac{p_F^3}{(m_1 \Delta_1)^{3/2}} \ll 1,
869: \end{equation}
870: while $R_g$ is still quite small, the contribution of doped carriers to $T_c$ through the Bloembergen-Rowland mechanism
871: can be neglected. Thus, if one exchange process dominates the physics are short distances, the short-range
872: exchange contribution has the same form, and can only differ in sign.
873: In what follows, we consider the general form of exchange integral Eq.(\ref{generic}), assuming that the exchange
874: constant $V_0$, its sign, and its range $R_0$ are those of the dominant indirect exchange contribution.
875: Note again that in our definition of exchange integral we sum over impurities twice.
876: We will also consider the case when the short-range part in Eq.(\ref{generic}) is rapidly oscillating,
877: \begin{equation}
878: J_{SR} = - V_0 \frac{R_0}{R}\, e^{-R/R_0} \cos{(\bm{Q}\cdot\bm{R})}.
879: \label{Osc}
880: \end{equation}
881: As we have seen above, this happens when two bands separated by wavevector $\bm{Q}$ contribute
882: the most to the short-range exchange interaction. $V_0$ has a negative sign for the Bloembergen-Rowland
883: contribution, and a positive sign for the superexchange.
884: When more than one process is important for short-range physics,
885: the problem can always be solved numerically for a given set of parameters.
886: Here we obtain an analytical solution in several limiting cases.
887:
888: \vglue 0.5cm
889:
890: \noindent
891: \textit{(1) Dilute system with almost no carriers.}
892:
893: \vglue 0.5cm
894:
895: In the absence of carriers, the type of order and $T_c$ is determined by the short-range part of
896: interaction. If ferromagnetic Bloembergen-Rowland interaction dominates at short distances, Curie temperature
897: is approximately given by the ferromagnetic interaction taken at the average distance between impurities
898: \cite{Korenblit}:
899: \begin{equation}
900: T_c \simeq 2.3 V_0 S^2 R_0 n_i^{1/3} \, e^{- \frac{0.87}{R_0 n_i^{1/3}}} \ \ V_0 > 0,
901: \label{perc}
902: \end{equation}
903: which is valid when $n_i \ll 1/R_0^3$. A comparison with Eq.(\ref{Curie}) shows that this
904: gives the following condition on the number of carriers:
905: \begin{eqnarray}
906: \label{cond}
907: p_F &=& (3 \pi^3 n_e)^{1/3} \ll p_{0F} \\
908: p_{0F}&=&\frac{27.6 V_0 \pi^2 S R_0}{(J^{pd}_{\bm{q}=0})^2 (S+1) n_i^{2/3}}\, e^{-0.87/(R_0 n_i^{1/3})} \nonumber
909: \end{eqnarray}
910: When the short-range interaction is antiferromagnetic, it favors spin glass order, and
911: ferromagnetism is absent when there are no or almost no carriers.
912: The condition on the number of carriers for ferromagnetism to be absent is then given by:
913: \begin{equation}
914: p_F \ll p_{0SG}=\frac{27.6 |V_0| \pi^2 R_0}{(J^{pd}_{\bm{q}=0})^2 n_i^{2/3}}\, e^{-0.87/(R_0 n_i^{1/3})}
915: \label{cond1}
916: \end{equation}
917: Finally, for a rapidly oscillating short-range part Eq.(\ref{Osc}) this condition can be rewritten as:
918: \begin{equation}
919: p_F \ll p_{0SG}=\frac{27.6 |V_0| \pi^2 R_0}{(S+1)(J^{pd}_{\bm{q}=0})^2 n_i^{2/3}}\, e^{-0.87/(R_0 n_i^{1/3})}
920: \label{cond2}
921: \end{equation}
922:
923: For the Anderson Hamiltonian, $J^{pd}_{q=0} = 2 V_1^2 a^3/\Delta_1$. Note that if the ratio $n_e/n_i$ is fixed,
924: $p_F$ grows much slower with $n_i$ than $p_0$. Thus, conditions in Eqs.(\ref{cond}),(\ref{cond1}),(\ref{cond2})
925: are likely to be satisfied at some finite concentration of magnetic impurities, $n_c \ll n_i \ll 1/R_0^3$.
926: For example, for antiferromagnetic or oscillating interactions, Eqs (\ref{cond1}),(\ref{cond2}) will define the concentation
927: of impurities above which carrier-driven ferromagnetism disappears.
928:
929: \vglue 0.5cm
930:
931: \noindent
932: \textit{(2) Dilute system with carriers.}
933:
934: \vglue 0.5cm
935:
936: When the carrier concentration is large, i.e., $p_F \gg p_0$ in Eqs.(\ref{cond}),(\ref{cond1}),(\ref{cond2}),
937: but the concentration of magnetic impurities is still small, $n_i \ll 1/R_0^3$, short-range interactions
938: between magnetic impurities will result in a correction to $T_c$, which
939: can be calculated by virial expansion (see, for example, Ref.\cite{BA}). Following Ref.\cite{AG},
940: since the range of RKKY interaction is large, we can represent $p-d$ interaction between carriers
941: and magnetic impurities by a mean field Zener Hamiltonian:
942: \begin{equation}
943: \hat{H}_{MF} = - J^{pd}_{\bm{q}=0} \bm{s} \sum_i \bm{S}_i,
944: \label{zener}
945: \end{equation}
946: where $\bm{s}$ is the density of ordered spin of the carriers, which is assumed to be constant
947: in space. In addition, there is a relatively short-range exchange interaction between impurity spins,
948: given by the processes described in Sections IV and V:
949: \begin{equation}
950: \hat{H}_{exch} = - \sum_{ij} J(\bm{R}_i - \bm{R}_j) \bm{S}_i \bm{S}_j.
951: \label{exx}
952: \end{equation}
953: The short-range exchange integral does not include the carrier contribution, since it is already
954: accounted for in the mean field Hamiltonian.
955: In the Zener model Eq.(\ref{zener}) , Curie temperature Eq.(\ref{Curie}) can be found by minimizing the free energy
956: density of the system of carriers and spins\cite{AG},
957: \begin{equation}
958: F = F_e + F_i,
959: \end{equation}
960: with respect to $\bm{s}$, and finding when the solution at small $s$ first appears. Here
961: \begin{equation}
962: F_e= \frac{(2 \mu_B s)^2}{2 \chi_0},
963: \end{equation}
964: and $F_i$ is given by the usual Zeeman term,
965: \begin{equation}
966: F_i = - n_i T \ln{\frac{\sinh{[J^{pd}_{\bm{q}=0} s (S + 1/2)/T]}}{\sinh{[J^{pd}_{\bm{q}=0} s/(2 T)]}}\,}
967: \end{equation}
968:
969: To find the virial correction to Eq.(\ref{Curie}) from Eq.(\ref{exx}), we need to include into $F$ the
970: contribution from two magnetic impurities, when they are close enough:
971: \begin{equation}
972: F_{2i} = \frac{1}{2}\, n_i^2 \int d^3 \bm{R} [F_{2i}(\bm{R}) - 2 F_i (\bm{R})],
973: \end{equation}
974: and calculate it from the Hamiltonian
975: \begin{equation}
976: \hat{H} = \hat{H}_{MF} + \hat{H}_{exch}.
977: \end{equation}
978: This can be done, since the Zeeman Hamiltonian for two spins, interacting via direct
979: exchange interaction, can easily be solved. The integral over this solution, however,
980: can only be taken with logarithmic accuracy. The contribution from two impurities is important
981: when the distance between them is
982: \begin{equation}
983: R \le R_0 \ln{\frac{|V_0|/T}{\ln{(|V_0|/T)}}}.
984: \end{equation}
985: Finding $F_{2i}$, and repeating the minimization over $s$, we obtain, as expected, that a for ferromagnetic exchange
986: interaction ($V_0 > 0$) Curie temperature is enhanced:
987: \begin{equation}
988: \frac{\delta T_c[F]}{T_c} \simeq \frac{4 \pi S}{3 (S + 1)}\, n_i R_0^3 \ln^3{\frac{|V_0|/T_c}{\ln{(|V_0|/T_c)}}}.
989: \label{ferr}
990: \end{equation}
991: For an antiferromagnetic exchange interaction ($V_0 < 0$), the Curie temperature is reduced:
992: \begin{equation}
993: \frac{\delta T_c[A]}{T_c} \simeq - \frac{4 \pi}{3}\, n_i R_0^3 \ln^3{\frac{|V_0|/T_c}{\ln{(|V_0|/T_c)}}}.
994: \label{aferr}
995: \end{equation}
996: A rapidly oscillating exchange interaction Eq(\ref{Osc}) gives:
997: \begin{equation}
998: \frac{\delta T_c[O]}{T_c} \simeq - \frac{4 \pi}{3 (S+1)}\, n_i R_0^3 \ln^3{\frac{|V_0|/T_c}{\ln{(|V_0|/T_c)}}},
999: \label{oscill}
1000: \end{equation}
1001: which is independent of the sign of $V_0$.
1002: \vglue 0.5cm
1003:
1004: \noindent
1005: \textit{(3) ``Dense'' system.}
1006:
1007: \vglue 0.5cm
1008:
1009: The most interesting situation is the case of a ``dense'' system of magnetic impurities,
1010: when $n_i \gg (m_1 \Delta_1)^{3/2}$. Since the short-range part of the interaction could
1011: have a range much larger than the lattice spacing, it need not be really dense.
1012: This requirement can be rewritten as $n_i \gg (\Delta_1 / D)^{3/2}$. In a dense system,
1013: the wave functions of the neighboring impurities overlap strongly, and the main exchange contribution
1014: arises as a result of this overlap, not the RKKY interaction through free carriers. Then
1015: ferromagnetism arises even when no carriers are present, if the Bloembergen-Rowland exchange process
1016: dominates the physics at short distances ($V_0>0$):
1017: \begin{equation}
1018: T_c = \frac{2 S (S+1) n_i}{3}\, J_{\bm{q}=0} = \frac{8 \pi S (S+1) n_i}{3} V_0 R_0^3.
1019: \end{equation}
1020: Note that $T_c$ does not depend on the carrier
1021: concentration, and should be much higher than that resulting from the RKKY
1022: interaction.
1023:
1024: The Bloembergen-Rowland exchange integral may be weaker than superexchange. In that case,
1025: ferromagnetism in a ``dense'' system is suppressed, and spin glass order is favored. This is always
1026: the case for indirect band gap
1027: semiconductors, where the short-range ferromagnetic exchange is absent.
1028:
1029: The Bloembergen-Rowland mechanism in case of direct band gap
1030: necessarily leads to an increase of maximum $T_c$ as a
1031: function of concentration of magnetic impurities $n_i$.
1032:
1033: \section{Application to GaAs:Mn.}
1034:
1035: Application of the Anderson model to a real system, such as Ga$_{1-x}$Mn$_x$As is somewhat more
1036: involved than the model that we considered above, since Mn ion is in 3d$^5$ configuration with a
1037: spin $S=5/2$. This configuration has 5 d-orbitals. For symmetry reasons,
1038: there may be more than just one conduction or valence band,
1039: which is the case for GaMnAs. Then, as we discussed above in Section V, one has to take into account all pairwise
1040: contributions to the exchange integral, Eqs.(\ref{exchch}),(\ref{exch}). In particular,
1041: there may be unusual contributions to RKKY,
1042: such as those considered in Section V.
1043: In general, one has also to sum over all orbitals in the Anderson Hamiltonian, not just spins,
1044: and take the Hund's rule, spin orbit, and crystal field splitting into account. Different orbitals
1045: may have different $V-s$ with conduction and
1046: valence bands. The result of this treatment, however, produces the same exchange integrals $J(R)$; for
1047: example, in case when spin-orbit and crystal field splitting is neglected, the result will still be
1048: given by the exchange integral Eqs.(\ref{exchch}),(\ref{exch}), with
1049: $|V_1|^2$ is replaced by $\sum_m |V_{1m}|^2$, and similarly for $|V_2|^2$ ,where $V_{1m}$ is
1050: $V_1$ for $m$-th orbital. The relation between corrections to energy levels and exchange integrals
1051: is different in this more realistic case. However, the integrals involved are the same, and, as
1052: we have seen above in section III, corrections to energy levels are rather small. Taking Hund's rule into
1053: account results in replacing $S=1/2$ operator
1054: in equations of the previous section by $\bm{S}/(2S)$, with $S = 5/2$. Thus, this leads to
1055: the same results as in the previous section, with somewhat redefined $V-s$. The relationship
1056: between the energy shift and $J_{pd}$ will change, but the
1057: actual form of $J(R)$ is determined by the energy spectrum only. Instead of $V_{\alpha m}$, we may introduce
1058: \begin{equation}
1059: |V_{\alpha}|^2 \equiv \frac{\sum_m |V_{\alpha m}|^2}{4 S^2}\, ,
1060: \end{equation}
1061: and use the form of exchange integrals that we have obtained in the previous sections (with $S=5/2$).
1062: Equivalently, this would mean that the expression for $J(\bm{R})$ in terms of $J_{pd}$-s (or $J_{RKKY}$) for conduction
1063: and valence bands and level position will stay the same as in the previous sections.
1064:
1065: For a $p$-type semiconductor, such as Ga$_{1-x}$Mn$_x$As, we can adopt an inverted picture,
1066: where the ``empty'' band is now the valence band (empty of holes), while the filled band is the conduction band.
1067: The Anderson Hamiltonian is then easily rewritten in terms of holes. There are two types of holes in GaAs -
1068: the heavy hole and the light hole. Since their masses are very different ($0.081 m_e$ and $0.51 m_e$),
1069: the main RKKY exchange contribution is produced by the heavy hole. On the other hand, the superexchange contribution
1070: from the light hole band has a much larger range, and thus could potentially be more important than the superexchange contribution
1071: from the heavy hole band, or the mixed superexchange contribution. However, as we have seen in previous sections, most short range
1072: contributions for a direct band gap semiconductor take the form of the second term in Eq.(\ref{generic}), although
1073: there may be some variations. The problem is that the amplitude of superexchange $V_0 \propto V^4/D^3 \propto m^3 $, where $D$ is
1074: the band width! So, for light bands the effective range of the interaction is large, but the payback is that the amplitude turns
1075: out to be small. An easy estimate for the light hole band in GaAs shows that for $V$-s of the order of $1 eV$ this band plays
1076: no role in ferromagnetism. The hole mass in the split-off band
1077: is $0.15 m_e$. This band could also play an important role in the superexchange interaction, although, once again, the
1078: amplitude for realistic parameters turns out to be extremely small. The electron mass is $m_{ge} = 0.063 m_e$
1079: for the main $\Gamma$-valley. The masses and gaps for $L$- and $X$- valleys are much larger, so we don't expect them
1080: to play much role. Thus, we arrive at a simplified picture, where only the
1081: heavy hole band and $\Gamma$-valley electrons are relevant.
1082: The CFR $Mn$ $d^6/d^5$ level,
1083: which is important for our analysis, is in the conduction band, $\Delta_1 =1.5 eV$ above the top of the valence band.
1084: We can see from Eqs (\ref{rc1rc0}),(\ref{sesese}) that, for this particular level position, the ferromagnetic Bloembergen-Rowland
1085: interaction has the range $R_0 \equiv R_{BR} \simeq \hbar/\sqrt{2 m_{hh} \Delta_1}$, while the range of the superexchange is
1086: $R_{se} \simeq R_{BR}/2$. The amplitude of the Bloembergen-Rowland term, $V_{BR}/V_{se} \simeq V_2^2 m_{ge}/(2 V_1^2 m_{hh})$,
1087: could become comparable or exceed the amplitude of heavy hole superexchange.
1088: Note that, in case of strong short-range ferromagnetic
1089: interaction, it would be energetically favorable for $Mn$ impurities to form ferromagnetic clusters.
1090: This, in turn, would reduce the Curie temperature. Clustering of $Mn$ impurities would make magnetic properties
1091: of this material crucially dependent on sample preparation.
1092: On the other hand, antiferromagnetic short-range interactions should be stronger at shorter distances,
1093: which would potentially lead to an exchange integral (in the absence
1094: of carriers), which changes its sign as a function of the distance between impurities. In general, for the particular
1095: situation when the $Mn$ level is almost at the bottom of conduction band, the range of antiferromagnetic superexchange
1096: is approximately $R_0/2$, and we may represent the total short-range exchange integral in the following form:
1097: \begin{equation}
1098: J(R) \simeq V_0 \frac{R_0}{R}\, [- \alpha \exp{(- 2 R/R_0)} + \exp{(- R/R_0)}].
1099: \label{GaAsex}
1100: \end{equation}
1101: Here $V_0 = V_{BR} > 0$, while $\alpha \simeq m_{hh} V_1^2/(m_{ge} V_2^2)$ is the ratio of superexchange and Bloembergen-Rowland
1102: amplitudes (up to a factor of 2), which depends on the hybridization of
1103: impurity d-level with the valence band $(V_2)$, of impurity d-level with the heavy hole band, $V_{hh}$, and the corresponding effective
1104: masses. When $\alpha > 1$ (which is likely the case here, since $m_{hh} \gg m_{ge}$), the exchange becomes
1105: antiferromagnetic at short distances for $R < R_0 \ln{\alpha}$. The virial correction to $T_c$
1106: for such exchange integral (assuming $T_c$ is determined mostly by the $p-d$ interaction of magnetic impurities and heavy holes)
1107: is then given by:
1108: \begin{widetext}
1109: \begin{equation}
1110: \frac{\delta T_c}{T_c} \simeq \frac{4 \pi S}{3 (S + 1)}\, n_i R_0^3 (\ln^3{\frac{V_0/T_c}{\ln{(V_0/T_c)}}} - [2 + (1/S)]\ln^3{\alpha}),
1111: \end{equation}
1112: \label{deltatc}
1113: \end{widetext}
1114: and could change sign as well, at some large doping level. If carriers are not present, this exchange integral alone, for $\alpha > 1$,
1115: would lead to a saturation or decrease of $T_c$ at large doping. On the other hand, when $\alpha < 1$, ferromagnetism
1116: gets significantly enhanced at short distances.
1117:
1118: In general, the interplay between various short range contributions leads to a rather complicated physics
1119: at short distances. While a detailed calculation requires precise knowledge of all hybridization parameters
1120: from the quantum chemistry, we can estimate the $Mn$ concentration at which the short distance
1121: physics becomes important by requiring $\delta T_c / T_c \sim 0.5$ in Eq.(\ref{deltatc})
1122: We take the estimate of $J_{pd} \sim 150 eV \AA^3$ and $T_c \sim 110K$ from Ref.\cite{Ohno}, $\Delta_1 \sim 1.5 eV$,
1123: and assume that $\alpha \simeq 1$. Then $V_0 = J_{pd}^2 m_{hh}^3/(8 \pi^2)$, and we get:
1124: \begin{equation}
1125: x_i \sim \left(\frac{a}{2 R_0 \ln{(V_0/T_c)}}\, \right)^3 \sim 8\%
1126: \end{equation}
1127:
1128:
1129: \section{Effects of disorder and interactions.}
1130:
1131: In this section we consider rather briefly effects of disorder and interactions. Since the superexchange and the Bloembergen-Rowland
1132: exchange interaction are governed by high-energy virtual processes, they are independent of disorder. The RKKY interaction,
1133: however, gets modified. This modification was first considered by de Gennes\cite{deGennes}, who argued that, since the
1134: RKKY interaction at large distances is dominated by the $2 p_F$ Kohn anomaly wavevector, the vertex corrections are
1135: not essential for the averages over disorder.
1136: The long-distance power law in the RKKY interaction then gets multiplied by an exponential factor $\exp{(- 2 R/l)}$,
1137: where $l$ is the scattering length. These effects were indeed taken into account by Ohno et al.\cite{Ohno} in their
1138: original paper. Abrahams \textit{et al.}\cite{elihu}, however, have shown, that this is not the whole story,
1139: since disorder introduces instead
1140: a \textit{distribution} of $J(R)$ at large distances. We note here that the long-distance behavior
1141: of the RKKY interaction is not
1142: essential for ferromagnetism. The Curie temperature is determined by RKKY exchange at \textit{short} distances,
1143: or $J_{RKKY}(\bm{q}=0)$. Of course, the vertex corrections are essential for the calculation of the RKKY loop diagram at
1144: $\bm{q} = 0$. Summing all ladder diagrams ($0$ order in $1/p_F l$), shown in Fig. \ref{fig7}, leads to a diffuson contribution,
1145: \begin{equation}
1146: \Pi(\bm{q}, \omega_n) = - \frac{\nu D \bm{q}^2}{|\omega_n| + D \bm{q}^2}\, ,
1147: \end{equation}
1148: which significantly modifies frequency dependence of RKKY at $\bm{q} = 0$. Here $D$ is the diffusion coefficient.
1149:
1150: \begin{figure}
1151: \includegraphics[width=3in]{fig7.eps}
1152: %\includegraphics[width=3in]{fig7.pdf}
1153: \caption{Impurity corrections to Curie temperature.}
1154: \label{fig7}
1155: \end{figure}
1156:
1157: However, the static ($\omega = 0$), not dynamic, part of the diagram in Fig. \ref{fig7}
1158: determines $T_c$, and it is not changed at all.
1159: Thus, to the leading order in $1/p_F l$ disorder does not modify the Curie temperature.
1160: The interactions, if not too strong, can also be included as the standard Fermi-liquid corrections to $\chi_0$ in Eq.(\ref{Curie}).
1161: Weak localization corrections (the Cooperon diagrams), however, should modify $T_c$. They can also be included in
1162: the same way as the standard weak localization corrections to spin susceptibility (see, for example, Ref.\cite{lee}),
1163: and Eq.(\ref{Curie}) should still be valid.
1164:
1165: Finally, we note that strong exchange interaction $J^{pd}$ could bind holes at $Mn$ sites, forming a shallow
1166: (or deep) complex magnetic impurity. This effect would
1167: reduce the hole concentration and the number of free $Mn$ spins, and thus lead to a reduction Curie temperature.
1168: The interactions between
1169: these complex magnetic impurities would be determined by the overlap of the corresponding wavefunctions.
1170:
1171: \section{Conclusions.}
1172:
1173: We have investigated the model of magnetic semiconductors in which magnetic impurities are treated in the
1174: framework of the Anderson model. We have shown that the effective Hamiltonian of this model is more rich than
1175: the usual p-d model considered in the literature. Effectively, in the Anderson model, wave functions of localized
1176: impurities develop a "tail", which could be long range. When the concentration of magnetic impurities is large
1177: enough, the overlap of wavefunctions on two different sites leads to a very strong exchange interaction, which
1178: is important, and could dominate the physics at Mn concentrations as low as 5\%. There are two contributions
1179: to this exchange interaction - superexchange, in which localized electrons are exchanged trough only one type of
1180: bands (either conduction or valence bands), and the Bloumbergen-Rowland term, when the exchange is through both
1181: conduction and valence bands. We have found that, in case of direct band gap, the
1182: Bloemberger-Rowland exchange is ferromagnetic. This could lead to a dramatic enhancement of Curie temperatures in
1183: certain magnetic semiconductors with a direct band gap, such as GaMnAs.
1184: One other important consequence of the
1185: indirect exchange is that, if it is ferromagnetic and reasonably long range, doped carriers are not necessary
1186: to mediate ferromagnetism. This leads potentially to a new class of high-temperature magnetic semiconductors,
1187: with high Curie temperatures determined entirely by the interaction between localized impurities,
1188: \textit{not Zener mechanism}. This emphasizes the effort to search
1189: for new materials, where ferromagnetism is not carrier-driven (for example, driven by the Bloembergen-Rowland mechanism).
1190: Another important consequence of the Anderson model is that, if there are more than
1191: one type of carriers (for example, light and heavy holes in GaAs), the long-range RKKY interaction becomes rather
1192: complicated, since it involves a ``mixed'' contribution. We have also found that, at large doping,
1193: the RKKY interaction for the Anderson model and the $p-d$ model is \textit{different} at short distances.
1194: The effective exchange interaction in the $U=\infty$ Anderson model for any FS is given by Eqs.(\ref{exchangeint})
1195: (\ref{exchch}),(\ref{exch}). A numerical solution of this effective Hamiltonian for a given set of parameters (determined
1196: from quantum chemistry) would give the answer for $T_c(n_i)$ in the most general case.
1197:
1198: We have briefly considered effects of disorder and interactions. We have shown that when De Gennes\cite{deGennes}
1199: approximation $p_F l \gg 1$ is applicable, disorder does not modify Curie temperature. Localization effects, however,
1200: do modify Curie temperature, although their effect could be reduced to removal carriers.
1201: Finally, in this paper we have not considered the effects of mixing of conduction and valence bands (such as kp). These effects
1202: should also be included in the full description.
1203: We should mention that application of the Anderson model to GaAs:Mn was also considered in Ref. \cite{Kikoin},
1204: although the limits of the Anderson model and their conclusions are different from ours.
1205:
1206: I would like to thank L. P. Gor'kov, Z. Fisk, and G. Kotliar for many useful discussions,
1207: and H. Weitering for sharing his experimental observations.
1208: I am also very greatful to Misha Zhitomirsky for pointing out Ref.\cite{Kikoin}.
1209:
1210: This work was supported by the University of Tennessee. I also acknowledge, with gratitute, the
1211: input on this work provided by the participants of the Aspen 2003 winter conference on "Complex Quantum Order",
1212: and Aspen Summer 2003 workshop on "Competing Orders and Quantum Criticality in Correlated Electrons, Bosons, and Spin Systems".
1213:
1214:
1215:
1216:
1217: \begin{thebibliography}{}
1218: \bibitem{Prinz} G.A. Prinz,
1219: Science, v.\textbf{282}, 1660 (1998).
1220: \bibitem{Wolf} S.A. Wolf, D.D. Awschalom, R.A. Buhrman, J.M. Daughton, S. von Moln\'ar, M.L. Roukes,
1221: A. Y. Chtchelkanova, and D.M. Treger,
1222: Science, v.\textbf{294}, 1488 (2001).
1223: \bibitem{Zener1}C. Zener,
1224: Phys. Rev. \textbf{81}, 440 (1951).
1225: \bibitem{Zener2}C. Zener,
1226: Phys. Rev. \textbf{83}, 299 (1951).
1227: \bibitem{AG} A. A. Abrikosov, L. P. Gor'kov,
1228: Zh. Exp. and Theor. Phys., \textbf{43}, 2230 (1962).
1229: \bibitem{Korenblit} I. Ya. Korenblit and E. F. Shender,
1230: Sov. Phys. Usp., v. \textbf{21}(10), 832 (1978) [Usp. Fiz. Nauk \textbf{126}, 233 (1978).]
1231: \bibitem{Furdyna} J. K. Furdyna and J. Kossut,
1232: ``Diluted Magnetic Semiconductors'', vol. \textbf{25} of ``Semiconductor and Semimetals''
1233: (Academic Press, New York, 1988).
1234: \bibitem{Dietll} T. Dietl,
1235: ``Diluted Magnetic Semiconductors'', vol. \textbf{3B} of ``Handbook of Semiconductors''
1236: (North-Holland, New York, 1994).
1237: \bibitem{dietl1} T. Dietl, H. Ohno, F. Matsukura, J.Cibert, D.Ferrand,
1238: Science, \textbf{287}, 1019
1239: \bibitem{dietl2}
1240: T. Dietl, H. Ohno, and F. Matsukura,
1241: Phys. Rev B \textbf{63}, 195205.
1242: \bibitem{Munekata} H. Munekata, H. Ohno, S. von Molnar, A. Segm\"uller, L. L. Chang, and L. Esaki,
1243: Phys. Rev. Lett. \textbf{63}, 1849 (1989).
1244: \bibitem{Ohno} F. Matsukura, H. Ohno, A. Shen, and Y. Sugawara,
1245: Phys. Rev. B \textbf{57}, R2037 (1998).
1246: \bibitem{deBoeck} J. De Boeck, R. Oesterholt, A. Van Esch, H. Bender, C. Bruynseraede, C. Van Hoof, and G. Borghs,
1247: Appl. Phys. Lett. \textbf{68}, 2744 (1996).
1248: \bibitem{Ohno1} H. Ohno, A. Shen, F. Matsukura, A. Oiwa, A. Endo, S. Katsumoto, and Y. Iye,
1249: Appl. Phys. Lett. \textbf{69}, 363 (1996).
1250: \bibitem{Ohno2} H. Ohno,
1251: Science \textbf{281}, 951 (1998).
1252: \bibitem{Park} Y. D. Park, A. T. Hanbicki, S. C. Erwin, C. S. Hellberg, J. M. Sullivan, J. E. Mattson,
1253: T. F. Ambrose, A. Wilson, G. Spanos, B. T. Jonker,
1254: Science \textbf{295}, 651 (2002).
1255: \bibitem{Millis} A. Chattopadhyay, S. Das Sarma, A. J. Millis,
1256: Phys. Rev. Lett. \textbf{87}, 227202 (2001).
1257: \bibitem{Larkin} V. G. Vaks, A. I. Larkin, S. A. Pikin,
1258: ZhETF \textbf{53}, 281 (1967); ZhETF \textbf{53}, 1089 (1967).
1259: \bibitem{Young} D.P. Young, D. Hall, M.E. Torelli, Z. Fisk, J.L. Sarrao, J.D. Thompson, H.R. Ott, S.B. Oseroff,
1260: R.G. Goodrich, R. Zysler,
1261: Nature, \textbf{397}(6718), 412 (1999).
1262: \bibitem{Akimitsu} J. Akimitsu, K. Takenawa, K. Suzuki, H. Harima , Y. Kuramoto,
1263: Science, \textbf{293} (5532), 1125 (2001).
1264: \bibitem{Fisk} Z. Fisk, H. R. Ott, V. Barzykin, and L. P. Gor'kov,
1265: Physica B \textbf{312-313}, 808 (2002).
1266: \bibitem{Sato} K. Matsubayashi, M. Maki, T. Tsuzuki, T. Nishioka, N. K. Sato,
1267: Nature, \textbf{420}, 143 (2002).
1268: \bibitem{SW} J. R. Schrieffer, P. A. Wolff, Phys. Rev. \textbf{149}, 491 (1966).
1269: \bibitem{BG1} V. Barzykin, L. P. Gor'kov,
1270: Phys. Rev. B \textbf{46}, 3059 (1992)
1271: \bibitem{GS1} L. P. Gor'kov, A. V. Sokol,
1272: J. Phys. France \textbf{50}, 2823 (1989); L. P. Gor'kov, A. V. Sokol,
1273: Pis'ma v ZheTF \textbf{48}, 505 (1988).
1274: \bibitem{BR} N. Bloembergen, T. J. Rowland,
1275: Phys. Rev \textbf{97}, 1679 (1955).
1276: \bibitem{GS2} L. P. Gor'kov, A. V. Sokol, Physica C \textbf{159}, 329 (1989).
1277: \bibitem{Keldysh} L. V. Keldysh,
1278: Zh. Exp.Theor.Phys., \textbf{45}, 364 (1963).
1279: \bibitem{BA} V. Barzykin, I. Affleck,
1280: Phys. Rev. B \textbf{61}, 6170 (2000).
1281: \bibitem{deGennes} P. J. de Gennes, J. Phys. Radium \textbf{23}, 630 (1962).
1282: \bibitem{elihu} A. Jagannathan, E. Abrahams, and M. J. Stephen,
1283: Phys. Rev. B \textbf{37}, 436 (1988).
1284: \bibitem{lee} P. A. Lee, T. V. Ramakrishnan, - Rev. Mod. Phys.\textbf{57}, 287 (1885).
1285: \bibitem{Kikoin} P. M. Krstaji\'c, V. A. Ivanov, F. M. Peeters,
1286: V. Fleurov, and K. Kikoin, Europhys. Lett., \textbf{61}, 235 (2003).
1287: \end{thebibliography}
1288: \end{document}
1289: