1: \documentclass[prl,twocolumn,showpacs]{revtex4}
2: %\documentclass[aps,showpacs]{revtex4}
3: %\documentclass[prb]{revtex4}% Physical Review B
4: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
5:
6: \usepackage{graphicx}
7: \usepackage{amsmath}
8: \usepackage{amssymb}
9:
10:
11: \begin{document}
12: \def\a{\alpha}
13: \def\b{\beta}
14: \def\e{\varepsilon}
15: \def\d{\delta}
16: \def\l{\lambda}
17: \def\m{\mu}
18: \def\t{\tau}
19: \def\n{\nu}
20: \def\o{\omega}
21: \def\r{\rho}
22: \def\S{\Sigma}
23: \def\G{\Gamma}
24: \def\D{\Delta}
25: \def\O{\Omega}
26: % Macro varie
27:
28: \def\ra{\rightarrow}
29: \def\ua{\uparrow}
30: \def\da{\downarrow}
31: \def\pd{\partial}
32: \def\bg{{\bf g}}
33: \def\br{{\bf r}}
34: \def\bm{{\bf m}}
35: \def\bz{{\bf z}}
36:
37: \def\be{\begin{equation}}
38: \def\ee{\end{equation}}
39: \def\bea{\begin{eqnarray}}
40: \def\eea{\end{eqnarray}}
41: \def\nn{\nonumber}
42: \def\lb{\label}
43: \def\pref#1{(\ref{#1})}
44:
45: \title{Origin of four-fold anisotropy in square lattices of circular
46: ferromagnetic dots}
47:
48:
49: \author{G.N.~Kakazei,$^{1,2}$ Yu.G.~Pogorelov,$^3$ M.D.~Costa,$^4$ T.~Mewes,$^{5}$
50: P.E.~Wigen,$^2$ P.C.~Hammel,$^2$ V.O.~Golub,$^{1}$ T.~Okuno,$^6$
51: V.~Novosad$^7$}
52:
53: \affiliation{$^1$Institute of Magnetism National Academy of Sciences
54: of Ukraine, 36b Vernadskogo Blvd., 03142 Kiev, Ukraine}
55:
56: \affiliation{$^2$Department of Physics, Ohio State University, 191
57: West Woodruff Avenue, Columbus, OH 43210}
58:
59: \affiliation{$^3$IFIMUP/Departamento de F\'{\i}sica, Universidade do
60: Porto, R. Campo Alegre, 687, Porto, 4169, Portugal}
61:
62: \affiliation{$^4$CFP/Departmento de F\'{\i}sica, Universidade do
63: Porto, R. Campo Alegre, 687, Porto, 4169, Portugal}
64:
65: \affiliation{$^5$MINT/Department of Physics and Astronomy,
66: University of Alabama, Box 870209, Tuscaloosa, AL 35487}
67:
68: \affiliation{$^6$Institute for Chemical Research, Kyoto University,
69: Kyoto 611-0011, Japan}
70:
71: \affiliation{$^7$Materials Science Division, Argonne National
72: Laboratory, Argonne, IL 60439}
73:
74: \begin{abstract}
75: We discuss the four-fold anisotropy of in-plane ferromagnetic
76: resonance (FMR) field $H_r$, found in a square lattice of circular
77: Permalloy dots when the interdot distance $a$ gets comparable to the
78: dot diameter $d$. The minimum $H_r$, along the lattice
79: $\langle11\rangle$ axes, and the maximum, along the
80: $\langle10\rangle$ axes, differ by $\sim$ 50 Oe at $a/d$ = 1.1. This
81: anisotropy, not expected in uniformly magnetized dots, is explained
82: by a non-uniform magnetization $\bm(\br)$ in a dot in response to
83: dipolar forces in the patterned magnetic structure. It is well
84: described by an iterative solution of a continuous variational
85: procedure.
86: \end{abstract}
87: \pacs{75.10.Hk; 75.30.Gw; 75.70.Cn; 76.50.+g}
88: \maketitle
89:
90: Magnetic nanostructures are of increasing interest for technological
91: applications, such as patterned recording media \cite{moser}, or
92: magnetic random access memories \cite{allwood}. One of the most
93: important issues for understanding their collective behavior is the
94: effect of long-range dipolar interactions between the dots
95: \cite{demokritov}. For the single-domain magnetic state of a dot,
96: the simplest approximation is that dots are uniformly magnetized and
97: interactions only define relative orientation of their magnetic
98: moments \cite{guslienko}. If so, the system of dipolar coupled dots
99: in a square lattice should be magnetically isotropic.
100:
101: However, in all known experimental studies of closely packed arrays
102: of circular dots, a four-fold anisotropy (FFA) was found, either by
103: Brillouin light scattering \cite{mathieu}, ferromagnetic resonance
104: (FMR) \cite{jung} or magnetization measurements (from hysteresis
105: loops) \cite{natali,zhu}. It is important to note that FFA exists in
106: both unsaturated samples and saturated ones (i.e. above vortex
107: annihilation point on the hysteresis loop). Hence it cannot be only
108: associated with vortex formation suggested in Ref. \cite{natali}. It
109: was instead qualitatively related to stray fields from unsaturated
110: parts of magnetization inside the dots \cite{mathieu}. However no
111: quantitative description of FFA in such systems was given up to now.
112: So the aim of this study is to explain quantitatively the deviations
113: from isotropy in terms of modified demagnetizing effect in a
114: patterned planar system at decreasing inter-dot distance, from the
115: limit of isolated dot to that of continuous film. The choice of
116: \emph{X}-band FMR techniques for this study has an advantage in
117: eliminating possible interference from domain (vortex) structure
118: \cite{kakazei03}. The variational theoretical analysis is followed
119: by micromagnetic simulations.
120:
121: Permalloy (Py) dots were fabricated with electron beam lithography
122: and lift-off techniques, as explained elsewhere \cite{novosad}. The
123: dots of thickness $t$ = 50 nm and diameter $d$ = 1 $\mu$m were
124: arranged into square arrays with the lattice parameter $a$ (center
125: to center distance) varying from 1.1 $\mu$m to 2.5 $\mu$m. The
126: dimensions were confirmed by atomic force microscopy and scanning
127: electron microscopy. Room temperature FMR studies were performed at
128: 9.8 GHz using a standard \emph{X}-band spectrometer. The dependence
129: of the FMR field $H_r$ on the azimuthal angle $\varphi_H$ of applied
130: field $\mathbf{H}$ with respect to the lattice [10] axis for almost
131: uncoupled dots ($a$ = 2.5 $\mu$m) is shown in Fig. \ref{fig1}a. Only
132: a weak uniaxial anisotropy of $H_r(\varphi_H)$ is present here,
133: which can be fitted by the simple formula $H_{r}(\varphi_H) =
134: H_{r,av} + H_{2}\cos2\varphi_H$. For the $a$ = 2.5 $\mu$m sample, we
135: found the average peak position $H_{r,av} \approx 1.13$ kOe and the
136: uniaxial anisotropy field $H_{2} \approx 5$ Oe. The latter value
137: remains the same for the rest of our samples, so this uniaxial
138: anisotropy is most probably caused by some technological factors.
139:
140:
141: \begin{figure}
142: \includegraphics[bb=100bp 20bp 220bp 235bp, scale=.7]{Fig1.eps}
143: \caption{In-plane FMR field in square lattices of 1 $\mu$m circular
144: Py dots as a function of field angle $\varphi_H$. a) The data for
145: lattice parameter $a = 2.5 \mu$m are well fitted by uniaxial
146: anisotropy (solid line). b) At $a = 1.1 \mu$m, the best fit (solid
147: line) is a superposition of FFA and uniaxial anisotropy (separately
148: shown by dashed line).} \lb{fig1}
149: \end{figure}
150:
151: With decreasing distance $a$ between dots, two changes are observed
152: in the $H_r(\varphi_H)$ dependence. First, $H_{r,av}$ decreases to
153: $\approx 1.09$ kOe at $a$ = 1.1 $\mu$m (Fig. \ref{fig1}b). Second, a
154: four-fold anisotropy (FFA) is detected in the samples with
155: $a\leq$1.5 $\mu$m by pronounced minima of $H_{r}(\varphi_H)$ at
156: $\varphi_H$ close to the lattice $\langle11\rangle$ axes. This
157: behavior is fitted by $H_r(a,\varphi_H) = H_{r,av}(a) + H_4(a)
158: \cos4\varphi_H + H_2 \cos2\varphi_H$, as shown in Fig. \ref{fig1}b.
159: The interdot distance dependence of $H_{r,av}$ and FFA field $H_{4}$
160: is shown in Fig. \ref{fig2}. Also such anisotropy is detected in the
161: FMR linewidth, smaller for $\langle11\rangle$ than for
162: $\langle10\rangle$ case (reaching $\sim$30\% at $a/d$ = 1.1).
163:
164: \begin{figure}
165: \includegraphics[bb=20bp 10bp 180bp 120bp, scale=1.3]{Fig2.eps}
166: \caption{ a) Average FMR field $H_{r,av}$ and FFA field $H_{4}$ as
167: functions of interdot spacing $a$. The points are the experimental
168: data and the solid lines present the 1st iteration theory (the
169: limits mark $H_r$ of isolated dot and continuous film). b) The same
170: data plotted against $(d/a)^2$ for $H_{r,av}$ and $(d/a)^4$ for
171: $H_{4}$ give excellent linear fits (dashed lines).} \lb{fig2}
172: \end{figure}
173:
174: The FFA effect, which could not arise in uniformly in-plane
175: magnetized cylindrical dots, is evidently related to a non-uniform
176: distribution of the magnetization $\bm(r,\varphi,z)$ (in cylindric
177: coordinates $0 \leq r \leq R = d/2, 0 \leq \varphi < 2\pi, 0\leq z
178: \leq t$). A similar effect was discovered using Brillouin light
179: scattering \cite{mathieu} and magnetization reversal
180: \cite{natali,zhu} in such systems under weak enough external fields,
181: which displace vortices in each dot. This can be modeled by
182: displacements of two oppositely in-plane magnetized uniform domains
183: \cite{guslienko01}. But in the presence of external fields strong
184: enough to observe FMR, one has to assume a continuous (and mostly
185: slight) deformation of $\mathbf m(r,\varphi)$. The simplest model
186: for such deformation uses a variational procedure with respect to a
187: single parameter \cite{metlov}. However, as will be shown below, the
188: non-uniform magnetic ground state of this coupled periodic system
189: results from a rather complicate interplay between intra-dot and
190: inter-dot dipolar forces, which requires a more general variational
191: procedure.
192:
193: Assuming fully planar and $z$-independent dot magnetization with the
194: 2D Fourier amplitudes $\bm_\bg = \int \textrm{e}^{i\bg \cdot \br}
195: \bm(\br)d\br$, the total (Zeeman plus dipolar) magnetic energy (per
196: unit thickness of a dot) can be written as (see Appendix)
197:
198: \be
199: \lb{eq1}
200: E = -\mathbf{H} \cdot \bm_0 +
201: \frac{2 \pi} {a^2} \sum_{\bg\neq 0} \frac{f(g t)}{g^2} |\bm_\bg
202: \cdot \bg|^2, \ee
203:
204: \noindent where $f(u)=1-\left(1-\textrm{e}^{-u}\right)/u$
205: \cite{guslienko} and the vectors of the 2D reciprocal lattice are
206: $\varphi_H$-dependent: $\bg = (2\pi/a) (n_1 \cos\varphi_H - n_2 \sin
207: \varphi_H, n_1 \sin \varphi_H + n_2 \cos \varphi_H)$ (for
208: $\mathbf{H}\parallel x$ and integer $n_{1,2}$). The variation of
209: exchange energy at deformations on the scale of whole sample is of
210: the order of stiffness constant ($\sim 10^{-6} \textrm{ erg/cm}$ for
211: Py) and it can be neglected beside the variation $\sim H M_s d^2
212: \sim 10^{-2} \textrm{ erg/cm}$ of terms included in Eq. \ref{eq1}.
213: If the dot magnetization has constant absolute value: $\bm (\br) =
214: M_s \left (\cos \varphi(\br), \sin \varphi(\br) \right)$, its
215: variation: $\d \bm (\br) = \hat{\bz} \times \bm (\br)
216: \d\varphi(\br)$ (where $\hat{\bz}$ is unit vector normal to plane),
217: is only due to the angle variation $\d \varphi (\br)$. Using the
218: Fourier transform $\d\bm_\bg = \hat{\mathbf{z}} \times
219: \sum_{\bg^\prime} \bm_{\bg - \bg^\prime}\d\varphi_{\bg^\prime}$ in
220: the condition $\d E = 0$ leads to the equilibrium equation for the
221: Fourier amplitudes:
222:
223: \bea
224: m_{\bg,y} & = & \frac{4\pi}{H a^2} \sum_{\bg^\prime \neq 0}
225: \frac{f(g^\prime t)} {{g^\prime}^2}
226: (\bm_{\bg^\prime} \cdot \bg^\prime) \notag \\
227: & & \qquad\qquad\qquad (\bm_{\bg - \bg^\prime} \times \bg^\prime)
228: \cdot \hat{\bz}. \lb{eq2} \eea
229:
230: \noindent It can be suitably solved by iterations:
231:
232: \bea
233: m^{(n)}_{\bg,y} &=& \frac{4\pi}{H
234: a^2}\sum_{\mathbf{g}^\prime\neq 0} \frac{f(g^\prime
235: t)}{{g^\prime}^2} (\bm^{(n-1)}_{\bg^\prime} \cdot
236: \bg^\prime)\nn \\
237: & & \,\qquad\qquad (\bm^{(n-1)}_{\bg - \bg^\prime}
238: \times\bg^\prime)\cdot\hat{\bz}, \lb{eq3} \eea
239:
240: \noindent starting from uniformly magnetized dots as zeroth
241: iteration: $m^{(0)}_{\bg,y} = 0$, $m^{(0)}_{\bg,x} = 2\pi R M_s
242: J_1(g R)/g$ (with the Bessel function $J_1$). Already the 1st
243: iteration (including the inverse Fourier transform):
244:
245: \bea
246: {m}^{(1)}_y(\br)& =& -\theta(d-2r) \frac{8\pi^2 R^2 M^2_s}{H
247: a^2} \sum_{\bg \neq 0} \frac{f(g t) g_x g_y}{g^3}\,\nn \\
248: & &\qquad\qquad\qquad\qquad J_1(g R) \cos(\bg\cdot\br), \lb{eq5}
249: \eea
250:
251: \noindent (with the Heavyside $\theta$ function) reveals the FFA
252: behavior, due to the rotationally non-invariant product $g_x g_y $.
253: The calculated maximum variation of $\varphi(\br) = \arcsin [m_y
254: (\br)/M_s]$ in $\langle10\rangle$ field geometry is $\sim60\%$
255: bigger than in the $\langle11\rangle$ geometry (Fig. \ref{fig3},
256: upper row). This expected behavior persists upon further iterations.
257: Our analytic approach was checked, using the micromagnetic OOMMF
258: code \cite{donahue} on a 9$\times$9 array of considered disks (Fig.
259: \ref{fig3}, lower row) at standard values of $M_{s} = 0.83$ kOe and
260: exchange stiffness $1.3 \cdot 10^{-6}$ erg/cm \cite{kakazei04} for
261: Py. The distributions obtained in this way for the central disk in
262: the array are within $\sim10\%$ to the analytic results of the 1st
263: iteration.
264:
265: \begin{figure}
266: \includegraphics[width=8cm]{Fig3.eps}
267: \end{figure}
268: \begin{figure}
269: \includegraphics[width=8cm]{Fig4.eps}
270: \caption{ Density plots of the equilibrium magnetization angle
271: $\varphi(\mathbf r)$ for two field geometries, $\varphi_H = 0$
272: ($\langle10\rangle$) and $\varphi_H = \pi/4$ ($\langle11\rangle$).
273: Upper row: calculated from the sum, Eq. \ref{eq5}, over
274: 100$\times$100 sites of reciprocal lattice at parameter values $a =
275: 1.1\, \mu$m, $H=$ 1.1 kOe, $M_s=$ 0.83 kOe. Lower row: micromagnetic
276: calculation by OOMMF code for the central disc in the 9$\times$9
277: array.} \lb{fig3}
278: \end{figure}
279:
280: The FMR precession of $\bm(\br)$ is defined by the internal field
281: $\mathbf{H}_i(\br) = \mathbf{H} + \mathbf{h}(\br)$ through the local
282: dipolar field
283:
284: \bea h_z(\br)&=&-\frac{4\pi}{a^2}\sum_{\bg}\left[1-f(g
285: t)\right]m_{\bg,z}\cos\left(\bg\cdot\br\right),\nn\\
286: h_\a(\br)&=&-\frac{4\pi}{a^2}\sum_{\b,\bg\neq 0} \frac{f(g t)g_\a
287: g_\b}{g^2}\tilde{m}_{\bg,\b}\cos\left(\bg\cdot\br\right), \lb{eq6}
288: \eea
289:
290: \noindent ($\a,\b=x,y$). The 1st iteration for $\mathbf{h}(\br)$
291: corresponds to the zeroth iteration for $\bm(\br) =
292: (M_s,\m_y,\m_z)$, which now includes the uniform FMR amplitudes
293: $\m_y,\,\m_z$. Then the local demagnetizing factors $N_x(\br) = -
294: h_x(\br) / M_s,\,N_{y,z}(\br) = -h_{y,z}(\br)/\m_{y,z}$ define the
295: local FMR field $H_r(\br)$:
296:
297: \bea
298: H_r(\br) &=& \sqrt{H_0^2+M_s^2[N_z(\br)-N_y(\br)]^2/4}\nn \\
299: &-& M_s[N_z(\br)+N_y(\br)-2N_x(\br)]/2\lb{eq7}
300: \eea
301:
302: \noindent (here $H_{0}\approx 3.3$ kOe). The average FMR field is
303: defined by the isotropic averaged demagnetizing factors
304:
305: \begin{eqnarray}
306: \overline{N}_{x,y}& = &\left(\pi R^2 \right)^{-1} \int_{r < R}
307: N_{x,y}
308: \left(\br\right)d\br \nonumber\\
309: &=& (8\pi^2/a^2)\sum_{\bg\neq 0}f(g t)J_1^2(g R)/g^2, \label{eq8}
310: \end{eqnarray}
311:
312: \noindent and $\overline{N}_z = 4\pi-2\overline{N}_x$. At
313: $a\to\infty$, they tend to the single dot values \cite{joseph} which
314: are for $t/R=0.1$: $N_{x,y}^{(d)}\approx0.776$ and $N_z^{(d)}
315: \approx 11.01$. Using $N_i^{(d)}$ instead of $N_i(\br)$ in Eq.
316: \ref{eq7} accurately reproduces the single dot FMR limit
317: $H_{r}^{(d)} \approx 1.15$ kOe (estimated from Fig. \ref{fig2}b).
318: Otherwise, for decreasing interdot distance, $a\to d$, the 1st
319: iteration values, Eq. \ref{eq8}, used in Eq. \ref{eq7} well describe
320: the tendency of $H_{r,av}(a)$ towards the continuous film limit
321: $H_r^{(f)}=\sqrt{H_0^2+4\pi^2M_s^2}-2\pi M_s\approx0.96$ kOe (Fig.
322: \ref{fig2}a).
323:
324: Finally, by calculating the true local FMR fields $H_{r}(\br)$ from
325: Eqs. \ref{eq6} and \ref{eq7}, the field dependent absorption is
326: obtained as $I(H)\propto\int_{r<R} \d(H-H_{r}(\br))d\br$. Then the
327: FMR fields, $H_r$ defined from maximum of $I(H)$ in two geometries,
328: display FFA in a good agreement with the experimental data (Fig.
329: \ref{fig2}). This effect is due to the fact that stronger
330: deformation of magnetization stronger suppresses the demagnetizing
331: effect (the differences $N_z-N_{x,y}$) and thus enhances $H_r$. Also
332: it produces a bigger spread of local resonance fields $H_r(\br)$ and
333: thus broadens the FMR line, again in agreement with our
334: observations.
335:
336: In conclusion, it is shown that under in-plane magnetic fields,
337: $\mathbf{H}$, even strong enough for FMR, the dipolar coupling in a
338: dense lattice of circular magnetic dots is able to produce a
339: continuous deformation of the dot magnetization, strongest for the
340: field orientation along lattice axes.
341:
342: Work at ANL was supported by the U.S. Department of Energy, BES
343: Materials Sciences under Contract No. W-31-109-ENG-38; MDC was
344: supported by FCT (Portugal) and the European Union, through POCTI
345: (QCA III) grant No. SFRH/BD/7003/2001.
346:
347:
348: \begin{thebibliography}{10}
349: \bibitem{moser}A. Moser, K. Takano, D.T. Margulies, M. Albrecht, Y. Sonobe,
350: Y. Ikeda, S. Sun, and E.E Fullerton, J. Phys. D: Applied Physics
351: \textbf{35}, R157 (2002); S. Sun, D. Weller, J. Magn. Soc. Jpn. 25,
352: 1434 (2001).
353: \bibitem{allwood}D.A. Allwood, Gang Xiong, M.D. Cooke, C.C. Faulkner,
354: D. Atkinson, N. Vernier, and R. P. Cowburn, Science \textbf{296},
355: 2003 (2002).
356: \bibitem{demokritov}S.O. Demokritov, B. Hillebrands, and A.N. Slavin,
357: Phys. Rep. \textbf{348}, 441 (2001).
358: \bibitem{guslienko}K.Yu. Guslienko and A.N. Slavin, J. Appl. Phys. \textbf{87}, 6337 (2000);
359: J. Magn. Magn. Mat. \textbf{215}, 576 (2000).
360: \bibitem{mathieu}C. Mathieu, C. Hartmann, M. Bauer, O. B\"{u}ttner, S. Riedling,
361: B. Roos, S.O. Demokritov, and B. Hillebrands, Appl. Phys. Lett.
362: \textbf{70}, 2912 (1997).
363: \bibitem{jung} S. Jung, B. Watkins, L. DeLong, J.B. Ketterson, and
364: V. Chandrasekhar, Phys. Rev. B \textbf{66}, 132401 (2002).
365: \bibitem{natali} M. Natali, A. Lebib, Y. Chen, I.L. Prejbeanu, and K. Ounadjela,
366: J. Appl. Phys. \textbf{91}, 7041 (2002).
367: \bibitem{zhu}X. Zhu, P. Grutter, V. Metlushko, and B. Ilic, Appl. Phys. Lett.
368: \textbf{80}, 4789 (2002).
369: \bibitem{kakazei03}G.N. Kakazei, P.E. Wigen, K.Y. Guslienko, R.W. Chantrell,
370: N.A. Lesnik, V. Metlushko, H. Shima, K. Fukamichi, Y. Otani, and V.
371: Novosad, J. Appl. Phys. \textbf{93}, 8418 (2003).
372: \bibitem{novosad} V. Novosad, K. Yu. Guslienko, H. Shima, Y. Otani, S. G. Kim,
373: K. Fukamichi, N. Kikuchi, O. Kitakami, and Y. Shimada, Phys. Rev. B
374: 65, 060402 (2002).
375: \bibitem{guslienko01}K.Yu. Guslienko, Phys. Lett. \textbf{278}, 293 (2001).
376: \bibitem{metlov} K.L. Metlov, Phys. Stat. Sol. (a) \textbf{189}, 1015 (2002);
377: K.L. Metlov and K.Yu. Guslienko, Phys. Rev. B \textbf{70}, 052406
378: (2004).
379: \bibitem{donahue}M.J. Donahue and D.G. Porter, URL: http://math.nist.gov/oommf
380: \bibitem{kakazei04}G.N. Kakazei, P.E. Wigen, K.Y. Guslienko, V. Novosad,
381: A.N. Slavin, V.O. Golub, N.A. Lesnik, and Y. Otani, Appl. Phys.
382: Lett. \textbf{85}, 443 (2004).
383: \bibitem{joseph}R.I. Joseph and E. Schl\"{o}mann, J. Appl. Phys. \textbf{36}, 1579 (1965).
384:
385: \end{thebibliography}
386:
387: \section{Appendix}
388: For fully planar and \emph{z}-independent dot magnetization, the
389: dipolar energy per unit thickness of a dot in the lattice is:
390:
391: \bea
392: E_d & = & \frac 1 {2t} \int_{-t/2}^{t/2} dz \int_{-t/2}^{t/2}
393: dz^\prime \int_c d\br \int d\br^\prime \sum_{\a,\b} m_\a(\br)\nn\\
394: & \times& \frac{\pd^2}{\pd r_\a \pd r_\b}\frac {m_\b(\br^\prime)}
395: {\sqrt{\left|\br - \br^\prime\right|^2 + \left(z -
396: z^\prime\right)^2}},
397: \nn
398: \eea
399:
400: \noindent where the 2D integrations $\int_c d\br$ and $\int d\br$
401: are respectively over the unit cell and over the entire plane. It
402: can be also presented as
403:
404: \bea E_d & = & \frac 1 {2t} \int_{-t/2}^{t/2} dz \int_c d\br
405: \sum_{\a} m_\a(\br) h_\a(\br,z) \nn\\
406: & =& \frac 1 {4\pi t a^2} \int_{-t/2}^{t/2} dz
407: \int_{-\infty}^{\infty} dq {\rm e}^{-i q z} \sum_{\a,\bg} \tilde
408: m_{\a,\bg} \tilde h_{\a,\bg,q},
409: \nn
410: \eea
411:
412: \noindent where the Fourier amplitudes of the dipolar field are:
413:
414: \bea
415: h_{\a,\bg,q} & = & \int_c d\br \int_{-\infty}^{\infty} dz^\prime
416: {\rm e}^{i (\bg \cdot \br + qz^\prime)} h_\a(\br,z^\prime) \nn\\
417: & = & \int_c d\br \int_{-\infty}^{\infty} dz^\prime {\rm e}^{i (\bg \cdot
418: \br + q z^\prime)} \int d\br^\prime \int_{-t/2}^{t/2} dz^{\prime\prime} \nn\\
419: & \times &
420: \sum_\b \frac{\pd^2}{\pd r_\a \pd r_\b} \frac {m_\b(\br^\prime)}
421: {\sqrt{\left|\br - \br^\prime\right|^2 + \left(z^\prime -
422: z^{\prime\prime}\right)^2}}.
423: \nn
424: \eea
425:
426: \noindent To calculate them, we express the lattice magnetization
427: $m_\b(\br^\prime)$ through its Fourier amplitudes:
428:
429: \[m_\b(\br^\prime) = \frac 1 {a^2} \sum_{\bg^\prime} {\rm e}^{-i \bg^\prime
430: \cdot \br^\prime} m_{\b,\bg^\prime},\]
431:
432: \noindent and then introduce the factor ${\rm e}^{i(\bg^\prime \cdot \br -
433: qz^{\prime\prime})}$ into the integral in $d\br^\prime dz^{\prime}$,
434: and the compensating factor ${\rm e}^{-i(\bg^\prime \cdot \br -
435: qz^{\prime\prime})}$ into the integral in $d\br dz^{\prime\prime}$.
436: Then the spatial integrations in $E_d$ are done accordingly to the formulas:
437:
438: \bea
439: &&\int_c d\br {\rm e}^{i (\bg - \bg^\prime) \cdot \br} = a^2
440: \d_{\bg,\bg^\prime},\nn\\
441: &&\,\quad\int_{-t/2}^{t/2} dz \int_{-t/2}^{t/2}
442: dz^{\prime\prime}{\rm e}^{i q (z^{ \prime
443: \prime} - z)} = \frac 4 {q^2} \sin^2 \frac {qt} 2,\nn\\
444: && \qquad\quad \int_{-\infty}^{\infty} dz^\prime \int d\br^\prime {\rm
445: e}^{i\bg \cdot (\br - \br^\prime) + i q(z^\prime -
446: z^{\prime\prime})}\nn\\
447: && \qquad\qquad \times\frac{\pd^2}{\pd r_\a \pd r_\b} \frac 1
448: {{\sqrt{\left| \br
449: - \br^\prime\right|^2 + \left(z^\prime - z^{\prime\prime}\right)^2}}}\nn\\
450: & & \qquad\qquad\qquad\qquad\qquad = \frac {4 \pi g_\a g_\b}{g^2 +
451: q^2}. \nn \eea
452:
453: \noindent Finally, the momentum integration
454:
455: \[\int_{-\infty}^{\infty} \frac{\sin^2 (qt/2)} {q^2({g}^2 +
456: q^2)} dq = \frac{\pi t}{2 {g}^2} f(g t)\]
457:
458: \noindent leads to the result included in Eq. \ref{eq1}.
459:
460:
461: \end{document}
462: