1: \documentstyle[aps,epsfig,multicol]{revtex}
2: %documentstyle[aps,prl,multicol,fancyheadings,epsfig,amsbsy,amssymb,amstex]{revtex}
3:
4: \begin{document}
5: % \draft command makes pacs numbers print
6: \draft
7: %inerst title
8: \title{Itinerant Ferromagnetism for Mixed Valence Systems}
9: % repeat the \author\address pair as needed
10: \author{C. D. Batista,$^1$ J. Bon\v ca,$^2$ and J. E. Gubernatis$^1$}
11: \address{$^1$Center for Nonlinear Studies and Theoretical Division
12: Los Alamos National Laboratory, Los Alamos, NM 87545\\
13: $^2$ Department of Physics, FMF, University of Ljubljana and J.
14: Stefan Institute, Ljubljana, Slovenia}
15: \date{\today}
16: \maketitle
17: \begin{abstract}
18:
19: We introduce a novel mechanism for the unusual itinerant
20: ferromagnetism found in mixed valence systems like
21: Ce(Rh$_{1-x}$Ru$_x$)$_3$B$_2$ , La$_x$Ce$_{1-x}$Rh$_3$B$_2$ , US,
22: USe, and UTe.
23: %We solve the Periodic Anderson model by means of the Constrained
24: %Path Monte Carlo method and a mean field theory. To derive the
25: %experimental consequences of the new mechanism.
26: %Using the Constrained-Path Monte Carlo method and mean-field
27: %theory, we show this mechanism is modeled by the periodic Anderson
28: %model.
29: With it we can provide an explanation for the long-unexplained
30: large value of $T_c$ ($\sim$ 100$^\circ$K) value and the maximum
31: in the magnetization below $T_c$ found experimentally.
32: %, and to complete
33: %%the characterization of this phase and confirmation of the
34: %mechanism we propose experiments.
35: We also show that this novel
36: itinerent ferromagnetism can be continuously connected with the
37: localized case for which the energy scale is much smaller
38: ($J_{RKKY} \sim$ 1$^\circ$K).
39:
40:
41: \end{abstract}
42: % insert suggested PACS numbers in braces on next line
43: \pacs{}
44:
45: %\vspace*{-0.4cm}
46: \begin{multicols}{2}
47:
48: \columnseprule 0pt
49:
50: \narrowtext
51: %\vspace*{-0.5cm}
52: % body of paper here
53: %\section{Introduction}
54: {\it Introduction.} Understanding the mechanisms for
55: ferromagnetism usually involves a reconciliation of a localized
56: electron picture, traced back to Heisenberg \cite{heisenberg}, and
57: an itinerant electron picture, traced back to Bloch \cite{bloch}.
58: For many insulating materials, adding itinerant features, such as
59: indirect exchange, to the localized picture brings satisfactory
60: agreement with basic experimental features. For many metals,
61: adding localized features, such as spin waves, to the itinerant
62: picture has a similar effect. On the other hand it is now
63: appreciated that for some novel classes of materials, such as the
64: heavy Fermion and mixed valence materials, doctoring one picture
65: or the other is often a questionable procedure. The electrons can
66: display both localized and itinerant properties as these materials
67: are neither good insulators nor good metals.
68:
69: In this letter we propose a simple mechanism for ferromagnetism
70: (FM) in heavy fermion mixed valence materials. The mechanism
71: relies on several specific energy scales set by the band structure
72: of these materials even though the electron correlation energy is
73: the dominant energy scale. The simplicity and generality of the
74: mechanism however should provide a useful framework for discussing
75: ferromagnetism in broad classes of heavy fermion mixed valence
76: materials often found in 4$f$ and 5$f$ materials and their compounds.
77: In fact we will show that it seems to explain some of the very
78: unusual long unexplained ferromagnetic properties of such Ce based
79: compounds as Ce(Rh$_{1-x}$Ru$_x$)$_3$B$_2$ \cite{malik}
80: and the uranian monochalcogenides \cite{santini}.
81:
82: The mechanism is based upon a two-band structure. The lower band
83: near the zone center is dispersive and nearly free of electron
84: correlation effects. Away from the zone center, this band is very
85: flat and doubly occupied states experience strong correlation
86: effects. The upper band is separated from the lower band by a gap
87: that is smaller than the correlation energy. When the chemical
88: potential is at the flat band level, the system becomes unstable
89: to the formation of a ferromagnetic state. There are compounds
90: satisfying this mixed valence requirement under normal conditions
91: or by either adding additional electrons (doping), applying
92: pressure, or changing the temperature. This ferromagnetic state is
93: a manifestation of a Hund-like rule among electrons in band states as
94: opposed to usual case of electrons in localized states (atomic
95: orbitals).
96:
97: We will show that the periodic Anderson model (PAM) is a simple
98: microscopic realization of this picture. Using quantum Monte Carlo
99: simulations, we will demonstrate that the model admits a
100: ferromagnetic ground state in the mixed valence regime with
101: electron occupancies reflecting our physical picture. We then
102: develop a mean field picture that reproduces the ground state
103: properties of the PAM and also allows us to compute finite
104: temperature properties. Under certain conditions we find the
105: finite temperature mean-field approximation predicts the same highly unusual
106: behavior as a peak, below $T_c$, in the temperature dependence of
107: the magnetization \cite{malik,santini} and the deviation of
108: the inverse susceptibility from Curie-Weiss behavior \cite{cheche}.
109:
110: {\it Model Hamiltonian.} The PAM is described by the Hamiltonian
111: \begin{eqnarray}
112: H &=& -t\sum_{\langle i,j \rangle,\sigma} (d_{i\sigma}^\dagger
113: d^{}_{j\sigma}+d_{j\sigma}^\dagger d^{}_{i\sigma})
114: +V\sum_{i,\sigma} (d_{i\sigma}^\dagger
115: f^{}_{i\sigma}+f_{i\sigma}^\dagger d^{}_{i\sigma}) \nonumber \\ & &
116: \quad\quad +\epsilon_f\sum_{i,\sigma}n_{i\sigma}^f +\frac{U}{2}
117: \sum_{i,\sigma}n_{i\sigma}^fn_{i\bar {\sigma}}^f
118: \label{eq:pam}
119: \end{eqnarray}
120: where $d_{i\sigma}^\dagger$ and $f_{i\sigma}^\dagger$ create an
121: electron with spin $\sigma$ in $d$ and $f$ orbitals at lattice
122: site $i$ and $n^f_{i\sigma}=f^{\dagger}_{i\sigma}f^{}_{i\sigma}$
123: is the number operator for the $f$-electrons of spin ${\sigma}$ at
124: site $i$. The hopping amplitude $t$
125: between $d$-orbitals is only to nearest-neighbor sites. The
126: hopping amplitude $V$ hybridizes different orbitals on the same
127: site.
128:
129: When $U=0$, the resulting non-interacting Hamiltonian $H_0$ has
130: two dispersive bands:
131: \begin{equation}
132: E_\sigma^\pm({\bf k})=\frac{1}{2} \Biggl[
133: e_{\bf k}+\epsilon_f \pm \sqrt{(e_{\bf k}
134: -\epsilon_f)^2+4V^2}
135: \Biggr]
136: \end{equation}
137: separated by an hybridization gap.
138: For a chain, the band structure is illustrated in Fig.~1a.
139: Irrespective of spatial dimension, each band has regions where the
140: ${\bf k}$-states are dominantly of $d$ or $f$-character. A mixed valence
141: condition arises when the Fermi energy $E_F$ sits in a crossover
142: region which is around the energy $\epsilon_f$. Three energies are
143: important: the hybridization gap $\Delta$, the $f$-state
144: dispersion $\delta_f=\partial E^{-}/\partial k|_{k_F}=\hbar v_F$ ($v_F$
145: is the Fermi velocity),
146: and the Coulomb (correlation) energy $U$.
147: $\delta_f$ is measuring the band dispersion close to the Fermi level. In the
148: interacting problem we will
149: assume that $U>\delta_f,\Delta$. As we shall now discuss, the
150: competition is between $\delta_f$ and $\Delta$ which will be a
151: competition between a ferromagnetic and paramagnetic state.
152:
153: {\it Mechanism for Ferromagnetism.} In Fig.~1a-c we illustrate the
154: physical mechanism for the ferromagnetic state which emerges when
155: the electron filling $\rho$ is arround 1/4
156: (one electron per lattice site)
157: and the energy $\epsilon_f$ of the $f$-orbitals is
158: close to $E_F$. The latter condition defines the intermediate
159: valence regime.
160:
161: To explain the mechanism, we first consider the non-interacting
162: case (Fig.~1a). States in the lower band with mainly $f$-character
163: have a small dispersion $\delta_f$ because of the absence of
164: direct hopping between the $f$-orbitals. In the interacting case,
165: because the Coulomb interaction only affects the electrons in the
166: $f$-orbitals, the electrons which were doubly occupying the states
167: in the lower band with mainly $d$-character are practically
168: unaffected. On the other hand, the electrons in the many states
169: which are close to the Fermi level and have mainly an
170: $f$-character will be strongly affected. These electrons spread to
171: higher unoccupied ${\bf k}$-states in the $f$-part of the band and
172: polarize by an in-band Hund's rule (Fig.~1b). To see the analogy
173: to Hund's rule more clearly, we consider the limiting case,
174: represented in Fig.~1c, where $\delta_f=0$ and $\Delta \neq 0$,
175: reducing Fig.~1b to a two level system in momentum space with each
176: level being strongly degenerate (we only plot the states with predominant $f$
177: character). If we add electrons to this two levels system
178: it is easy to show that the ferromagnetic solution has the lowest energy and is
179: therefore the ground state. By polarizing, the spatial part of the
180: wave function becomes antisymmetric and there is no double occupancy
181: in the real space. In this way, the Coulomb repulsion is reduced
182: to zero. The kinetic energy has also the lowest possible value if the electrons
183: are occupying the lower energy (degenerated) levels (see Fig.~1c).
184:
185:
186:
187: We now describe the conditions for the stability of the
188: ferromagnetic state, and we will see how this
189: mechanism departs from the one of Stoner \cite{stoner}.
190: To this end we state the condition $U \gg \Delta$. From Fig.~1b
191: it is clear that the cost of the spreading of the ferromagnetic
192: solution is an increase in the kinetic energy proportional to $\delta_f$.
193: On the other hand, if we build up a nonmagnetic solution only
194: using the states of the lower band, there will be a minimum delocalization
195: for each particle due to the fact that a finite set of $f$-character
196: states is in the upper band (see Fig.~1). This can be seen by
197: constructing an $f$ Wannier function with the lower band states.
198: This Wannier function will have a delocalization lenght which depends
199: on the ${\bf k}$ wave vector where the two bands are crossing.
200: Therefore, by localizing in real space to avoid the double occupancy,
201: the electrons will have a finite
202: probability of occupying the $f$-states in the upper band. The
203: energy cost per electron of occupancy is proportional to $\Delta$.
204: If the hybridization gap is much larger than the dispersion of the
205: states, that is, $\Delta \gg \delta_f$, the ferromagnetic state lies
206: lower in energy than the nonmagnetic state. The energy of the
207: excited nonmagnetic state is proportional to $\Delta$. This can
208: be easily seen from the limiting (and non-realistic) case $\delta_f=0$
209: ploted in Fig.~1c, for which the analogy with the atomic shells and
210: the Hund's rule is evident.
211:
212:
213: \begin{figure}[tbp]
214: \begin{center}
215: \vspace{-1.0cm}
216: \epsfig{file=Fig1ab.ps,width=70mm,angle=-0}
217: \end{center}
218: \end{figure}
219: \begin{figure}[tbp]
220: \begin{center}
221: \vspace{-3.0cm}
222: \epsfig{file=Fig1cprl.ps,width=68mm,angle=-0}
223: \vspace{-5.1cm}
224: \end{center}
225: \caption{Illustration of the ferromagnetic mechanism.}
226: \label{fig1}
227: \end{figure}
228:
229:
230: It is important to emphasize that the hybridization gap $\Delta$
231: and $f$-state dispersion $\delta_f$ are the two basic ingredients
232: of our mechanism: $\Delta$ only appears if there is more than one
233: band and $\delta_f$ is small only if the hybridization is weak. If
234: $U>\Delta,\delta_f$, then $\Delta$ is the energy scale of the
235: paramagnetic state, while $\delta_f$ is the energy scale of the
236: ferromagnetic solution. We can estimate the magnitude of these
237: scales for the PAM: for $V \lesssim t/2$, we get $\delta_f \sim
238: \partial e_k/\partial k|_{k_F} V^2/(e_{k_F}-\epsilon_f)^2$
239: and $\Delta \sim V^2/(2dt-\epsilon_f)+
240: \sqrt{(2dt+\epsilon_f)^2+4V^2}-2dt$. Therefore, if $\epsilon_f$ is
241: close to the bottom of the lower band ($\epsilon_f+2dt \sim t$)
242: and the Fermi level is close to the top ($E_F \sim \epsilon_f-V^2/(2dt-\epsilon_f)$),
243: $\Delta$ is considerably larger than $\delta_f$ and the ground
244: state is ferromagnetic. We can see that the ferromagnetic
245: solution is stable for comfortably realistic values of the
246: parameters. It is clear from this analysis the particle density $\rho$
247: for a ferromagnetic solution must be close to quarter filling.
248:
249: {\it Quantum Monte Carlo Method.} Using the constrained-path Monte
250: Carlo method (CPMC), we computed the ground-state properties of
251: the PAM on a square lattice. The CPMC method projects the ground
252: state from an initial state $|\psi_T\rangle$ by converting the
253: iterative procedure
254: \begin{equation}\label{eq:cpmc}
255: |\psi_{i+1}\rangle =e^{-\tau H}|\psi_{i}\rangle
256: \end{equation}
257: into a branched random walk. The details of the method are
258: described elsewhere \cite{zhang} as are our prior uses of it on
259: the square PAM \cite{bonca,batista}.
260:
261: The defining characteristic of the CPMC method is its elimination
262: of the fermion sign problem by excluding random walkers
263: $|\phi\rangle$ that violate $\langle \psi_T|\phi\rangle>0$. If
264: $|\psi_T\rangle$ were the exact ground state $|\psi_0\rangle$,
265: this constraint would generate an exact elimination of the sign
266: problem, and hence an exact solution. Extensive benchmarking
267: indicates the CPMC method provides accurate estimates of the
268: energy and various correlation functions.
269:
270: We prepared the trial state in specific values of the total spin
271: and z-component of spin \cite{batista}. Because the
272: Hamiltonian(\ref{eq:pam}) conserves these quantum numbers, the
273: iterative process (\ref{eq:cpmc}) produces a ground state with the
274: same $S$ and $S_z$. In the ferromagnetic regime, when the number of
275: lattice sites $N$ was greater than the number of electrons $N_e$,
276: the resulting energy as a function of $S$ almost always showed a
277: minimum at a value of $S$ satisfying
278: $0<S<\frac{1}{2}(2N-N_e)$. Hence we typically found a partially
279: polarized ferromagnetic ground state.
280:
281: For square lattices we also computed dependence of the static spin structure
282: factor and the electron occupancies both on position and wave
283: number space \cite{bonca}. Most notable for the present work is
284: the wave vector dependence of the different spin components of the
285: upper and lower band electron occupancies shown in Fig.~2. It has
286: the same quantitative features as Fig.~1b.
287:
288:
289: {\it Mean Field Theory.} The quantum Monte Carlo simulations are
290: limited to chains and square lattices of relatively small sizes at
291: zero temperature. However, we have found that their predictions
292: are well described by a simple spin-polarized mean-field
293: approximation. We can easily extend this approximation to larger
294: systems sizes in higher dimensions and at finite temperatures. We
295: can use large lattices in three-dimensions at finite temperatures
296: to compare with the results of experiments.
297:
298: %There are two reasons to expect mean field theory to give a good
299: %description of the ferromagnetic phase: One, the condition
300: %$U>\Delta>\delta_f$ for the ferromagnetic state is satisfied even
301: %if $U<t$ (weak coupling). Two, the ferromagnetic ground state
302: %suggested above is very similar to a partially polarized
303: %non-interacting eigenstate. (Of course the picture described in
304: %Fig.1 only depicts the essential features of the mechanism.)
305:
306: The mean field Hamiltonian $H_{MF}$ is
307: \begin{equation}
308: H_{MF} = H_{0} + \frac{U}{2}
309: \sum_{i,\sigma} (\langle n_{i\sigma}^f \rangle n_{i\bar
310: {\sigma}}^f + n_{i\sigma}^f \langle n_{i\bar {\sigma}}^f \rangle-
311: \langle n_{i\sigma}^f \rangle \langle n_{i\bar {\sigma}}^f
312: \rangle) \label{eq:mf}
313: \end{equation}
314: $\langle n_{i \uparrow}^f \rangle$ and $\langle n_{i \downarrow}^f
315: \rangle$ are the determined in a self-consistent way within the
316: mean-field theory. The translational symmetry of $H$ implies that
317: we can always find a translationally invariant ground state. This
318: means that $\langle n_{i\uparrow}^f \rangle$ and $\langle n_{i
319: \downarrow}^f \rangle$ will not depend on the site index $i$.
320: Therefore there are two variational parameters to be determined
321: from the mean-field equations: $\langle n_{\uparrow}^f \rangle$
322: and $\langle n_{\downarrow}^f \rangle$.
323:
324: Even though the mean field approximation can give a good
325: description of the FM ground state, it is well known that such
326: approximations overestimate the energy of the paramagnetic phase
327: because they improperly estimate the real space correlations that
328: are very important for the paramagnetic solution. Therefore it is
329: crucial to check that the ground state is the FM one by using a
330: more accurate method to evaluate the energy of the PM state. To
331: this end we calculated the ground state of the PAM with $6 \times
332: 6$ unit cells using the CPMC technique. We
333: found excellent agreement with the energy and as seen in Fig.~2
334: the electron occupancies.
335:
336: \begin{figure}[tbp]
337: \begin{center}
338: \vspace{-1.2cm}
339: \epsfig{file=Fig2prl.ps,width=80mm,angle=-90}
340: \vspace{-2.3cm}
341: \end{center}
342: \caption{Comparison between Mean Field and CPMC values of the
343: occupation numbers of the non-interacting band states
344: for each spin polarization in a $6\times6$ cluster
345: ($\rho=7/24$,$V=0.5t$, $U=t$, and $\epsilon_f=-3t$). On the
346: $x$-axis we only include the nonequivalent wave vectors ordered
347: according to increasing non-interacting energies.}
348: \label{fig2}
349: \end{figure}
350:
351: {\it Consequences.} Bolstered by this agreement we used the
352: mean-field approximation to calculate various thermodynamic
353: properties of the ferromagnetic phase for the three-dimensioanl
354: cubic PAM. In Fig.~3a we show the temperature dependence of the
355: magnetization for different values of $\epsilon_f$. From the
356: arguments given above it is clear that the magnetization comes
357: from the electrons which are occupying the states with
358: $f$-character. By increasing $\epsilon_f$ we are reducing the
359: number of $f$-electrons and therefore the magnitude of the zero
360: temperature magnetization. Some critical value of $\epsilon_f$
361: exists where the chemical potential starts departing from
362: $\epsilon_f$ and a new energy scale $\epsilon_f-E_F$ emerges in
363: the system. This new energy scale is reflected in the appearance
364: of a magnetization peak. (See the $\epsilon_f=-t$ case in
365: Fig.~3a.) To understand the last statement one first has to
366: realize that the zero temperature magnetization is small due to
367: the reduction of the number of $f$-electrons. When the temperature
368: is of the order of $\epsilon_f-E_F$, electrons are promoted
369: electrons from the double occupied band states to the
370: $f$-character states which have a large entropy (large density of
371: states). The $f$-electrons will be polarized due to the energy
372: considerations above explained. In this way we can explain the
373: origin of the magnetization peak below $T_c$. It is important to
374: note that the source for the large entropy is associated with
375: charge and not with spin degrees of freedom. This fact explains
376: why a state with larger magnetization has a higher entropy. The
377: magnetization curves shown in Fig. 3a are in good qualitative
378: agreement with the magnetization versus temperature data measured
379: in Ce(Rh$_{1-x}$Ru$_{x}$)$_3$B$_2$ for different values of $x$
380: \cite{malik}.
381:
382: % Brief description of $Ce(Rh$_{1-x}$Ru$_{x}$)$_3$B$_2$
383:
384: \begin{figure}[tbp]
385: \begin{center}
386: \vspace{-1.2cm}
387: \epsfig{file=Fig3prl.ps,width=90mm,angle=-90}
388: \vspace{-2.5cm}
389: \end{center}
390: \caption{Mean field results ($\rho=1/4$,$V=0.5t$, $U=2t$) for (a) magnetization,
391: (b) inverse susceptibility,
392: (c) density of states and (d) specific heat.}
393: \label{fig3}
394: \end{figure}
395:
396:
397: Another interesting aspect of this ferromagnetism is the deviation
398: from linearity (Curie-Weiss behavior)
399: for the inverse magnetic susceptibility above
400: $T_c$ over a large temperature range (see Fig. 3b) \cite{cheche}.
401: The behavior contrasts that predicted by theories based on
402: localized magnetic moments.
403:
404: From above it is also clear that the change in entropy from the
405: magnetic to the paramagnetic phase depends on the number for $f$
406: electrons (spin degrees of freedom). This dependence is seen in
407: Fig.~3d where we plot the calculated specific heat for different
408: values of $\epsilon_f$.
409:
410: A characteristic of this ferromagnetic state that has consequences
411: for the photoemission experiments is the implication that the
412: quasi-particle dispersion should be close to that of the electrons
413: in the non-interacting case. We recall that the ferromagnetic
414: state described in Fig. 1 is very similar to a non-interacting
415: polarized state. As in the Stoner mechanism, the main role of the
416: Coulomb repulsion is to polarize states with well defined
417: momentum. The hybridization gap of the noninteracting solution is
418: most likely replaced by a pseudogap (see Fig.~3c). This pseudogap should be seen
419: in the optical conductivity measurements.
420:
421:
422: {\it Conclusions.} Our quantum Monte Carlo simulations and
423: mean-field calculations clearly show the existence of an itinerant
424: ferromagnetic phase in mixed valence materials that is supported
425: by the PAM. This new phase describes qualitatively many experimental
426: features such as an unusually large value for $T_c$($\sim$
427: 100$^\circ$K) and the maximum in the magnetization below $T_c$
428: \cite{shaheen} that are found in such ferromagnetic compounds as
429: Ce(Rh$_{1-x}$Ru$_x$)$_3$B$_2$ \cite{malik},
430: Ce(Rh$_{1-x}$Os$_x$)$_3$B$_2$ \cite{malik}, and
431: La$_x$Ce$_{1-x}$Rh$_3$B$_2$ \cite{shaheen}.
432:
433: Many previous theories of 4$f$ and 5$f$ electron materials treated
434: these materials as systems of localized moments in the
435: $f$-orbitals. These theories are thus unable to describe an
436: itinerant ferromagnetic phase and experimental consequences
437: \cite{santini} like the peak in the magnetization below $T_c$
438: observed in some of these systems \cite{malik,cheche}, the large
439: value of $T_c$, the deviation of $\chi^{-1}(T)$ from the
440: Curie-Weiss law above $T_c$ in the uranium monochalcogenides
441: \cite{cheche}, and the mixed valence behavior of these compounds
442: \cite{shaheen,kanter,erbudak}.
443:
444:
445:
446: The physical picture just presented, when combined with our
447: previous results \cite{batista}, allows a reconciliation of the
448: localized and delocalized ferromagnetism pictures painted by
449: Heisenberg and Bloch seventy years ago in the sense that it is
450: possible to go continuously from the mixed valence FM state where
451: the $f$-electrons are delocalized to a FM state where there is one
452: localized electron in each $f$-orbital \cite{batista}. In our
453: picture one can do this by decreasing $\epsilon_f$ from the Fermi
454: level to a value near or below the bottom of the valence band. We
455: note that the energy scale of a localized ferromagnetic state is
456: that of the RKKY interaction, which according to de Gennes's rule
457: \cite{degennes} is of the order of 1$^\circ$K, while the scale for
458: the itinerant case introduced here is that of the hybridization
459: gap $\Delta$, which is of the order of 100$^\circ$K for heavy
460: fermion compounds \cite{note}. In going from the delocalized (mixed valence)
461: to the localized regime we thus expect a strong reduction of $T_c$
462: accompanied by an increase of the zero temperature magnetization.
463: (The $T=0$ magnetization is proportional to the $f$ occupation
464: number.) This expectation is consistent with the observed behavior
465: of the magnetization in La$_x$Ce$_{1-x}$Rh$_3$B$_2$ as function of
466: $x$ \cite{shaheen}.
467:
468:
469:
470: {\it Acknowledgements.} This work was sponsored by the US DOE
471: under contract W-7405-ENG-36. We acknowledge useful discussions
472: with A. J. Arko, B. Brandow, M. F. Hundley, J. J. Joyce,
473: J. M. Lawrence, S. Trugman, and J. L. Smith. In particular,
474: we thank J. M. Lawrence
475: for pointing out the experimental work on the Ce compounds.
J. B.
476: acknowledges the support Slovene Ministry of Education, Science and
477: Sport, Los Alamos National Laboratory and FERLIN.
478:
479:
480:
481:
482: % BIBLIOGRAPHY
483:
484: \begin{thebibliography}{99}
485:
486: \bibitem{heisenberg}
487: W. Heisenberg, Z. Physik {\bf 49}, 619 (198).
488:
489: \bibitem{bloch}
490: F. Bloch, Z. Physik {\bf 57}, 545 (1929).
491:
492: \bibitem{malik}
493: S. K. Malik, A. M. Umarji, G. k. Shenoy, and M. E. Reeves, Phys. Rev. B
494: {\bf 31}, 4728 (1985).
495:
496: \bibitem{santini}
497: P. Santini, R. L\'emanski, and P. Erdos, Advan. in Phys. {\bf 48}, 537
498: (1999).
499:
500: \bibitem{cheche}
501: V. I. Chechernikov, A. V. Pechennikov, E. I. Yarembash, L. F. Martinova,
502: and V. K. Slavyanskikh,
503: Sov. Phys. JETP {\bf 26}, 328 (1968).
504:
505:
506: \bibitem{stoner}
507: E. C. Stoner, Phil. Mag. {\bf 15},1018 (1933).
508:
509: \bibitem{zhang}
510: Shiwei Zhang, J. Carlson and J. E. Gubernatis, Phys. Rev.
511: Lett., {\bf 74}, 3652 (1995); Phys. Rev. B {\bf 55}, 7464 (1997);
512: J. Carlson, J. E. Gubernatis, G. Ortiz, and Shiwei Zhang, Phys.
513: Rev. B {\bf 59}, 12788 (1999).
514:
515: \bibitem{bonca}
516: J. Bon\v ca and J. E. Gubernatis, Phys. Rev. B {\bf 58}, 6992
517: (1998).
518:
519: \bibitem{batista} C. Batista, J. Bon\v ca, and J. E. Gubernatis,
520: Phys. Rev. B, to appear (cond-mat/0009128).
521:
522: %Experiments
523:
524:
525: \bibitem{shaheen}
526: S. A. Shaheen, J. S. Schilling, and R. N. Shelton, Phys. Rev. B {\bf 31},
527: 656 (1985).
528:
529: \bibitem{degennes} P.-G. de Gennes, Comm. Energie At. (France)
530: Rappt. {\bf 925}, (1959).
531:
532: \bibitem{note}
533: This change in the energy scale of both ferromagnetic states
534: is obtained with Monte Carlo by taking the energy difference between
535: the PM and the FM solutions at $T=0$. The mean field theory
536: is not appropiate to describe this transition because it overestimates
537: the PM energy.
538:
539: \bibitem{kanter}
540: M. A. Kanterand C. W. Kazmierowicz, J. Appl. Phys. {\bf 35}, 1053 (1954).
541:
542: \bibitem{erbudak}
543: M. Erbudak and J. Keller, Z. Phys. B {\bf 23}, 281 (1979).
544:
545: \end{thebibliography}
546:
547: \end{multicols}
548:
549: \end{document}
550:
551:
552: